ArticlePDF Available

Advancements in Electronic Materials and Devices for Stretchable Displays

Wiley
Advanced Materials Technologies
Authors:

Abstract and Figures

A stretchable display would be the ultimate form factor for the next generation of displays beyond the curved and foldable configurations that have enabled the commercialization of deformable electronic applications. However, because conventional active devices are very brittle and vulnerable to mechanical deformation, appropriate strategies must be developed from the material and structural points of view to achieve the desired mechanical stretchability without compromising electrical properties. In this regard, remarkable findings and achievements in stretchable active materials, geometrical designs, and integration enabling technologies for various types of stretchable electronic elements have been actively reported. This review covers the recent developments in advanced materials and feasible strategies for the realization of stretchable electronic devices for stretchable displays. In particular, representative strain‐engineering technologies for stretchable substrates, electrodes, and active devices are introduced. Various state‐of‐the‐art stretchable active devices such as thin‐film transistors and electroluminescent devices that consist of stretchable matrix displays are also presented. Finally, the future perspectives and challenges for stretchable active displays are discussed.
This content is subject to copyright. Terms and conditions apply.
2201067 (1 of 37) © 2023 Wiley-VCH GmbH
www.advmattechnol.de
Review
Advancements in Electronic Materials and Devices
for Stretchable Displays
Yeongjun Lee,* Hyeon Cho, Hyungsoo Yoon, Hyunbum Kang, Hyunjun Yoo,
Huanyu Zhou, Sujin Jeong, Gae Hwang Lee, Geonhee Kim, Gyeong-Tak Go, Jiseok Seo,
Tae-Woo Lee, Yongtaek Hong,* and Youngjun Yun*
DOI: 10.1002/admt.202201067
and exchange data with other devices
and systems over the internet. Further-
more, going beyond the connection of
devices, the “Internet of Things” (IoT)
has emerged, where everything around
us is connected.[1] Devices will evolve into
human-centric devices that require more
user-convenience, portability, and better
connectivity with their surroundings, and
people will demand wearable devicesthat
are easier to carry around.[2]
In recent years, there has been a pro-
liferation of wearable electronics such
as smart watches, clothes, and patches,
and even implantable devices to detect
physiological signals for healthcare moni-
toring. With the advent of wearable tech-
nology, the medical system is shifting
from managing a few patients in hospitals
to remotely monitoring several people who
have potential risks to their health during
daily activities. Although commercial
wristwatches are useful and widely used,
some people still do not want to wear them because they can be
uncomfortable and inaccurate. In particular, during a high level
of physical activity, they cannot obtain accurate data because of
losing skin contact or motion artifacts.[3–5] This is leading to the
development of stretchable and skin-like electronics,in which
electronic functions are introduced to the surface of the skin to
maximize a user’s comfort during daily life and ensure secure
contact with the skin to detect physiological signals.[6]
A stretchable display would be the ultimate form factor for the next generation
of displays beyond the curved and foldable configurations that have ena-
bled the commercialization of deformable electronic applications. However,
because conventional active devices are very brittle and vulnerable to mechan-
ical deformation, appropriate strategies must be developed from the material
and structural points of view to achieve the desired mechanical stretchability
without compromising electrical properties. In this regard, remarkable findings
and achievements in stretchable active materials, geometrical designs, and
integration enabling technologies for various types of stretchable electronic
elements have been actively reported. This review covers the recent devel-
opments in advanced materials and feasible strategies for the realization of
stretchable electronic devices for stretchable displays. In particular, representa-
tive strain-engineering technologies for stretchable substrates, electrodes,
and active devices are introduced. Various state-of-the-art stretchable active
devices such as thin-film transistors and electroluminescent devices that
consist of stretchable matrix displays are also presented. Finally, the future
perspectives and challenges for stretchable active displays are discussed.
Y. Lee
Department of Chemical Engineering
Stanford University
Stanford, CA 94305, USA
E-mail: yeongjunlee@gmail.com
H. Cho, H. Yoon, H. Yoo, S. Jeong, G. Kim, J. Seo, Y. Hong
Department of Electrical and Computer Engineering
Inter University Semiconductor Research Center (ISRC)
Seoul National University
Seoul 08826, Republic of Korea
E-mail: yongtaek@snu.ac.kr
The ORCID identification number(s) for the author(s) of this article
can be found under https://doi.org/10.1002/admt.202201067.
H. Kang, G. H. Lee, Y. Yun
Organic Material Lab.
Samsung Advanced Institute of Technology (SAIT)
Samsung Electronics
Suwon 16678, Republic of Korea
E-mail: youngjun.yun@samsung.com
H. Zhou, G.-T. Go, T.-W. Lee
Department of Materials Science and Engineering
Seoul National University
Seoul 08826, Republic of Korea
T.-W. Lee
School of Chemical and Biological Engineering
Institute of Engineering Research
Research Institute of Advanced Materials, Soft Foundry
Seoul National University
Seoul 08826, Republic of Korea
1. Introduction
Information and communication technology (ICT) has evolved
toward higher speed, density, and resolution, with lower power
consumption, and has ushered in an era of mobile devices such
as laptops, smart phones, and tablets. Interactions between
mobile devices have proliferated with advances in device per-
formance, and various services have been provided to connect
Adv. Mater. Technol. 2023, 2201067
2365709x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/admt.202201067 by Seoul National University, Wiley Online Library on [08/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com
© 2023 Wiley-VCH GmbH
2201067 (2 of 37)
www.advmattechnol.de
The display is the most important front-end interface for
communication between users and devices, and thus occupies
a fairly large area in the appearance of the device. This makes a
stretchable display essential for skin-like electronics (Figure 1).
Although the research on stretchable displays is still in the early
stage, much research is being actively reported. Display form
factor innovation is accelerating. Thus, stretchable displays are
expected to be commercialized in the near future.[7]
In the mid-2010s, flat-panel displays, which are usually less
than 10 cm thick, completely replaced cathode ray tube (CRT) dis-
plays, which were introduced in 1939 by the Radio Corporation
of America (RCA). It took almost 70 years to change the form
from heavy and cumbersome to a thin, light, and flat screen.
Flexible displays, which are designed to withstand being folded,
bent, and twisted, as opposed to the traditional flat-panel dis-
plays used in most devices, allow innovative smartphone designs
with various forms.[8] In the late 2010s, interest in various form
factors such as multi-foldable, rollable, and stretchable displays
began to increase with the release of foldable phones in the con-
sumer market. Although this market is still in an early adopter
phase, it is expected to be a mainstream in the near future. A
foldable display is a type of flexible display that can be folded and
unfolded like paper. This includes not only a single fold but also
multiple folds such as G-folding, in which it is folded inward
twice in a “G” shape, and Z-folding, in which it is folded in a “Z
shape, with only a third of the screen exposed to the outside. A
rollable display is also a type of flexible display with the ability to
expand the display area by pulling on it vertically or horizontally
so that the screen behaves like paper on a scroll.
Stretchable displays, which can be stretched in all directions
to change their shape, will be the next generation of displays
with a new form factor a step beyond that of flexible displays.
They can be deformed in any direction, freely folded, and fit
on any surface. Based on numerous studies of stretchable elec-
tronic materials, strain releasing geometries, stretchable back-
planes, and stretchable light-emitting diodes (LEDs), a few
prototypes of stretchable displays have been demonstrated. For
example, Samsung demonstrated a 9.1-inch stretchable active-
matrix organic light-emitting diode (AMOLED) display with 5%
stretchability at the Society for Information Display (SID) in
2017.[9] Someya etal. reported a highly elastic, 1 mm thick skin
display, which consisted of a 16 × 24 array of micro-light emit-
ting diodes (μ-LEDs) with 45% stretchability, at the American
Association for the Advancement of Science (AAAS) Annual
Meeting in Austin, Texas in 2018. In 2020, Beijing Oriental
Electronics (BOE) reported a stretchable AMOLED with a 10%
stretch ratio.[10] In 2021, Lee et al. showed the feasibility of a
skin-like display patch with a passive-matrix (PM) OLED array,[11]
and Royole demonstrated a 2.7 in. 96 × 60 pixel (42 pixels per
inch, ppi) μ-LED display with 120% stretchability at SID.[12]
Based on such interest and the research conducted in academia
and industry, it is expected that stretchable displays and skin-
like electronic devices will be realized in the near future and will
provide unprecedented types and shapes of electronic devices.
This reviewoutlines the recent progress made in the develop-
ment of deformable and stretchable displays.First, it discusses the
current technical approaches and prospects for strain engineering,
including pre-strained structures, island-bridge structures, and
intrinsically stretchable materials that could be implemented for
stretchable display arrays. Second, stretchable electrodes for the
interconnections are discussed, including intrinsically stretch-
able conductors (e.g., conductive nanomaterial composites, liquid
metals (LMs), and conductive polymers) and structural designs.
Third, the stretchable thin film transistors (TFTs) made from var-
ious semiconducting materials that are capable of being used for
a stretchable backplane are summarized. In particular, the results
of strain-engineered inorganic stretchable TFTs and intrinsically
stretchable TFTs that exploit semiconducting carbon nanotubes
(CNTs) and polymer semiconductors are discussed. Following
this, stretchable light-emitting devices such as inorganic, organic,
and metal-halide perovskites LEDs, along with alternating current
electroluminescence (ACEL) devices, are discussed. Finally, we
report the recent progress on stretchable display arrays and future
perspectives on high-density stretchable displays.
2. Structural Strategies for Stretchable Display
Most electronic materials such as semiconductors and conduc-
tors are brittle; however, they can be made more flexible by
decreasing their thickness. Downscaled materials can withstand
significantly larger strains and stresses than their bulk states.
However, they cannot aord a tensile strain that is greater
than the bending strain. To apply these electronic materials to
stretchable devices, there have been several approaches to make
electronic materials and devices stretchable at the macro-scale
level (extrinsically stretchable geometries) or at the micro-scale
level (intrinsically stretchable materials).
This section describes representative strategies for structural
designs that are viable for stretchable displays, including geo-
metrical engineering approaches to produce 1) wrinkled struc-
tures using a pre-strain method, 2) island-bridge structures,
and 3) approaches that use intrinsically stretchable materials
that exploit nanomaterial composites, molecular blending, and
molecular chemistry.
2.1. Prestrained Structures
The prestrain method is a prevalent method for forming a wrin-
kled structure for a flexible device using a pre-stretched elastomer
substrate. Wrinkled structures are flattened after restretching,
which allows the flexible devices to be stretched without
Figure 1. Conceptual image of stretchable displays in the potable elec-
tronics and skin-like wearables.
Adv. Mater. Technol. 2023, 2201067
2365709x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/admt.202201067 by Seoul National University, Wiley Online Library on [08/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com
© 2023 Wiley-VCH GmbH
2201067 (3 of 37)
www.advmattechnol.de
mechanical failure.[13,14] The wrinkled structure is fabricated by the
deposition or lamination of a flexible film on a pre-stretched elas-
tomeric substrate.[15] The first pre-strain-induced stretchable wrin-
kled structure was developed by the transfer of thin silicon ribbons
to a pre-strained elastomeric substrate (poly(dimethylsiloxane),
PDMS). After peeling o the silicon ribbons from the elasto-
meric substrate, they had a periodic wavy structure (Figure 2a).[16]
Because the prestrain method achieves reliable and stable
stretchable properties, various materials and devices have been
researched to make stretchable devices and systems.[17–22]
The theoretical model of a buckled structure has been sug-
gested to design the materials and geometries for stretchable
devices with the pre-strain method. The wavy configuration can
be determined by various factors such as the thickness, elastic
modulus, adhesion, Poisson ratio of the materials, and geom-
etry of the structure.[15] First, the wavelength (λ0) and amplitude
(A0) of the buckling are calculated by the model with a well-
defined sinusoidal wavy structure with a buckled geometry.[23,24]
In this model, the elastomeric substrate is assumed to be a very
thick film (semi-infinite solid), which was widely researched for
early wearable electronics:
,1
0
c
0
pre
c
h
Ah
λπ
ε
ε
ε
==
(1)
where h is the thickness of the sti material, εc is the critical
buckling strain, and εpre is the strain of the prestrained elasto-
meric substrate. However, this buckled geometry can be formed
only when εpre is larger than the critical buckling strain. The
critical buckling strain, εc, is determined by the elastic modulus
values of the elastomeric substrate, Es, and sti film, Ef:
1
4
3
c
s
f
2/3
E
E
ε
=
(2)
However, delamination can occur between sti materials and
soft elastomeric substrates. In this case, the model cannot pre-
dict the buckling behavior. Therefore, a new model has been
suggested to explain the finite deformation and non-linear
strain displacement relation.[15,25,26] Using this model, wave-
length λ and amplitude A of the buckling are calculated as
follows:
11
,
A
11
0
pre
1/3
0
pre
1/21/3
A
λλ
εξ
εξ
() ()
() ()
=++ =
++
(3)
where 5(
1)
32
pr
ep
re
ξεε
=
+
. In addition, the maximum strain of the
film, which is called the peak strain, can be calculated as the
sum of the membrane strain and bending strain in the buck-
ling structure. Peak strain εpeak is calculated as follows:
21
1
peak cpre
1/2
1/3
pre
1/2
εε
ε
ξ
ε
()
()
()
=
+
+
(4)
The value of εpre is much larger than that of εpeak. Therefore,
a pre-strained structure can be eectively used for a stretchable
Figure 2. Prestrain-induced wrinkled stretchable structure. a) Schematic process diagram of prestrain method. Reproduced with permission.[16] Copy-
right 2006, American Association for the Advancement of Science. b) Peak strain and membrane as a function of the pre-strain in Si ribbon/PDMS
substrate system. Reproduced with permission.[25] Copyright 2007, PNAS. c) Schematic diagrams of global and local buckling and local buckling.
Reproduced with permission.[27] Copyright 2008, AIP Publishing.
Adv. Mater. Technol. 2023, 2201067
2365709x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/admt.202201067 by Seoul National University, Wiley Online Library on [08/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com
© 2023 Wiley-VCH GmbH
2201067 (4 of 37)
www.advmattechnol.de
device (Figure 2b).[25] In addition, the maximum pre-strain
before fracture, εpre,max, can also be calculated by this model:
4143
144
pre,max
fracture
2
c
fracture
2
c
εε
ε
ε
ε
=+
(5)
where εfracture is the fracture strain of the sti material. There-
fore, the maximum prestrain is limited by the mechanical prop-
erties of the materials.
This well-defined wavy structure can only be formed on a
thick elastomeric substrate, which is considered a semi-infinite
solid. However, global buckling occurs instead of periodical
local buckling when the elastomeric substrate is thin. The
mode of buckling is illustrated in Figure2c.[27] To predict the
buckling mode, a pre-strained structure can be modeled with
rigidity. The critical global buckling strain can be calculated as
follows:
1
11.2
c,global
c
sf
c
F
Gh h
F
EA
ε
()
=
++
(6)
where
4/
c
22
FEIL
π
= is the critical buckling load,
is the eec-
tive shear modulus of a composite beam, hs is the thickness of
the substrate, hf is the thickness of a sti film, L is the length
of the composite beam, ss
ff
EA
Eh
Eh=+
is the eective ten-
sile rigidity, and ()4()
12
ff
2
ss
22
ffss sf
2
EI Eh Eh EhEh hh
EA
=−+ +
is the eective bending rigidity.[27] When εc,global is larger than
εc, local buckling occurs. However, if εc,global is smaller than εc,
global buckling occurs. To avoid this global buckling, the mate-
rials and geometrical structure must be designed appropriately
according to the model. Moreover, the adoption of a thin layer
between the film and substrate can eliminate global buck-
ling.[17] Using these theoretical models, the geometrical param-
eters and mechanical properties of the materials are the keys to
designing a stretchable system using pre-strain methods.[23,28]
2.2. Island-Bridge Structures
The structure engineering method of implementing a stretch-
able display is to divide it into a soft area and hard area on a
plane. The rigid electronic devices (e.g., inorganic/organic
LEDs and TFTs) are placed in island-shaped rigid pixelated
areas and connected through interconnects that are intrinsi-
cally or externally (geometrically) stretchable (Figure 3a). In
this structure, the devices in the rigid pixels are hardly aected
by external strain. Thus, this structure can impart mechanical
stretchability to well-developed rigid devices, including LEDs
and TFTs, which has the advantage of using the mature manu-
facturing process that has been established for rigid displays
over previous decades. Thus, the island–bridge structure has
been the most widely adopted method to implement stretchable
displays.[10,12,29,30]
One method of connecting rigid/flexible islands is to use
metallic bridges.[31] A serpentine-shaped bridge is mainly used
with metal thin film electrodes, where the stretching motion is
performed by 3D torsions of the flexible bridges.[32] Because the
local strain distribution and resultant distortion stress under
stretching are dependent on the geometry of the island–bridge
structures, the global strain range of the stretchable display is
modulated by geometrical factors such as the ratio between
island and bridge regions, design of the bridge interconnects,
in-plane/out-of-plane distortion, and buckling.[33–36]
The island and serpentine bridge structures can be fabri-
cated on one plane by patterning a silicon on insulator (SOI)
wafer[37,38] or plastic substrate such as polyimide (PI).[34,39] A
low island/bridge ratio can increase the elongation with a wide
deformation area, but there is a limit to the increase in resolu-
tion due to a small pixel area ratio. It is possible to improve
the elongation at a high island/bridge area ratio by applying
bridges with various curvatures, horseshoes, and fractal pat-
terns (Figure 3b).[40–43] In the serpentine structures, because
the strain is mainly concentrated on the inner and outer edges
of the curves,[44,45] it is necessary to limit the elongation, which
can minimize fatigue failure due to repeated stretching.
The rigid island and bridge interconnect regions can be
separately fabricated on the elastomeric substrate, and it is
also possible to apply not only serpentine flexible metal films
but also intrinsic stretchable conductors (e.g., LMs, conductive
nanocomposites, and conducting polymers).[11,46–48] The intrin-
sically stretchable interconnect can be fabricated in a straight
shape that minimizes the area occupied by the wavy flexible
interconnect, thereby improving the resolution. Furthermore,
intrinsically stretchable interconnects can be patterned in a
wavy shape to further improve the elongation and mechanical
stability.[49,50] In addition, the rigid island region can be located
above or below the elastomer or embedded therein to modulate
the strain distribution and provide more options in terms of
device structures and fabrication processes.[9,51]
The disconnection of interconnects near the edge of the
island pixel area due to localized strain can limit the stretch-
ability of the island-bridge structures.[52–54] To improve the
stretchability of the stretchable display array, various strain
engineering approaches that have applied lower Young’s mod-
ulus materials (Figure 3c),[55] sti materials with a gradient
modulus (Figure 3d),[56–61] or pillar structures (Figure 3e)[62]
have been reported. In addition, a kirigami-based design for a
stretchable display without image distortion was suggested.[63]
The μ-LEDs were transferred onto an electrical circuit board
with kirigami cutting patterns (Figure 3f). The meta-atom
expanded in the perpendicular direction to the loading with a
strain identical to the loading strain, resulting in a Poisson’s
ratio of –1 (Figure3g,h).
2.3. Intrinsically Stretchable Structures
Unlike the aforementioned structural methods for fabri-
cating stretchable electronics, intrinsically stretchable devices
would not require geometry-based strain engineering. Thus,
they would suciently increase the device stretchability while
avoiding the reduction of both the overall device density and
complexity of the device processes.[6466] Therefore, introducing
inherently stretchable materials (e.g., conductors and semi-
conductors) into stretchable electronic devices is an important
method to develop stretchable displays. Intrinsically stretchable
electronic materials are mostly based on polymeric materials
as elastic matrices or substrates that incorporate conducting
Adv. Mater. Technol. 2023, 2201067
2365709x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/admt.202201067 by Seoul National University, Wiley Online Library on [08/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com
© 2023 Wiley-VCH GmbH
2201067 (5 of 37)
www.advmattechnol.de
or semiconducting materials. Because polymer materials can
be functionalized with additional unique properties such as
self-healing, biocompatibility, and antioxidation properties,
intrinsically stretchable electronic materials could broaden the
functionality of skin-like electronics, including the develop-
ment of damage-resilient and biocompatible biometric devices
beyond those used for conventional applications.[67,68] This sec-
tion discusses recent representative approaches and their elec-
tronic device applications in terms of both inherent material
designs and physical blending approaches for improving the
stretchability while maintaining the electrical/optical properties.
Intrinsically stretchable conductors have been intensively
developed and utilized in stretchable electrodes and current
collectors to demonstrate intrinsically stretchable TFTs,[66,69,70]
LEDs,[64,71,72] photovoltaics (PVs),[73] and high-frequency
diodes.[74] Furthermore, intrinsically stretchable conductors
are used as the mechanically robust interconnects of stretch-
able device arrays. Various kinds of conducting materials such
as LMs, conducting polymers, and conductive nanomaterials
(e.g., Ag nanowires (AgNWs), Ag flakes, and CNTs) have been
widely used as intrinsically stretchable conductors. For exam-
ples, Matsuhisa et al. recently demonstrated the intrinsically
stretchable display and sensor systems (Figure 4a) by inte-
grating high-frequency (HF) diodes (Figure 4b), current col-
lectors, stretchable antennas, and strain sensors.[74] The con-
ducting polymer poly(3,4-ethylenedioxythiophene):polystyrene
Figure 3. Representative island-bridge structure. a) Stretchable LEDs with rigid islands and serpentine interconnects, along with b) their geometric
parameters. Reproduced with permission.[40] Copyright 2015, Elsevier. Various strain engineering approaches with c) lower Young’s modulus mate-
rials,[55] d) gradient Young’s modulus materials,[60] and e) pillar structures.[62] Reproduced with permission.[55] Copyright 2020, Wiley-VCH. Reproduced
with permission.[60] Copyright 2016, Wiley-VCH. Reproduced with permission.[62] Copyright 2020, American Chemical Society. f) Stretchable micro-LED
display designed with auxetic metamaterials. Digital images of g) a stretchable micro-LED meta-display with Poisson’s ratio of -1 and h) display attached
to a hemisphere. Reproduced with permission.[63] Copyright 2022, Wiley-VCH.
Adv. Mater. Technol. 2023, 2201067
2365709x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/admt.202201067 by Seoul National University, Wiley Online Library on [08/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com
© 2023 Wiley-VCH GmbH
2201067 (6 of 37)
www.advmattechnol.de
Figure 4. Representative material candidates and their applications to intrinsically stretchable electronics. Digital images of a) oxidized eGaIn-based
interconnects for stretchable wireless sensor and display system and b) high-frequency and intrinsically stretchable diodes. c) Optical microscope
image of AgNW-elastomer composite for stretchable current collectors. Reproduced with permission.[74] Copyright 2021, Springer Nature. d) Mole-
cular structure of PEDOT:PSS. Reproduced with permission.[64] Copyright 2022, Springer Nature. e) Schematic and optical microscope image of PR-
PEDOT:PSS-based conducting polymer for stretchable electrodes. Reproduced with permission.[64] Copyright 2022, Springer Nature. f ) Schematic and
g) crack onset strain with terpolymer approaches for modulating the mechanical properties of CPs. Reproduced with permission.[65] Copyright 2021,
Springer Nature. h) CP/SEBS elastomer blends for intrinsically stretchable TFTs. Reproduced with permission.[69] Copyright 2017, American Associa-
tion for the Advancement of Science. i) Digital images of intrinsically stretchable polymer light-emitting layers composed of CP/PU elastomer blends.
Reproduced with permission.[64] Copyright 2022, Springer Nature.
Adv. Mater. Technol. 2023, 2201067
2365709x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/admt.202201067 by Seoul National University, Wiley Online Library on [08/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com
© 2023 Wiley-VCH GmbH
2201067 (7 of 37)
www.advmattechnol.de
sulfonate (PEDOT:PSS) with ionic additives, AgNWs/tough
thermoplastic polyurethane (T-TPU)-based elastomeric com-
posites (Figure4c), oxidized eutectic gallium-indium (oxidized
eGaIn) printed on a stretchable substrate (Figure4a), and single
wall CNTs (SWCNTs) were used for high-frequency diodes,
current collectors, stretchable antennas, and strain sensors,
respectively. In addition, Zhang et al. reported highly e-
cient and intrinsically stretchable polymer LEDs (PLEDs) that
showed stable operation with up to 100% strain by developing
highly stretchable PEDOT:PSS electrodes with a polyrotaxane
(PR)-based photopatternable crosslinker (PR-PEDOT:PSS)
(Figure 4d,e).[64] Furthermore, Jiang et al. successfully demon-
strated intrinsically stretchable and biophysical sensor arrays
based on the aforementioned PR-PEDOT:PSS systems.[75]
Intrinsically stretchable semiconductors have also been
actively developed and utilized as the active layer for intrinsi-
cally stretchable devices. Among the several material candidates
for stretchable semiconductors, conjugated polymers (CPs)
have emerged as key materials because of their excellent optical
and electrical properties, with inherently lower Young’s moduli
compared to those of silicon and inorganic semiconductors.
However, the major challenge faced in developing CPs lies in
maintaining good electrical, optical, and mechanical proper-
ties simultaneously. Because of the extended π-conjugation of
the CP backbone, which is vital for electronic properties, they
often have rigid and semicrystalline properties. To lower the
rigidity and enhance the stretchability of CPs, a significant
structural modification process for CPs has been developed by
incorporating dynamic bonding units, conjugation breakers,
and several flexible chemical structures.[67,76–78] A random ter-
polymer approach was also eectively used to simultaneously
control the backbone rigidity and thin-film morphology. For
example, Mun etal. developed high mobility (>1 cm2 V1 s1)
and stretchable (>100% strain) diketopyrrolopyrrole (DPP)-
based CPs by incorporating two dierent types of randomly
distributed co-monomer units, which reduced the overall crys-
tallinity and long-range orders while maintaining short-range
ordered aggregates (Figure4f).[65] As a result, the resulting CPs
showed improved electrical and mechanical properties com-
pared to those of a control system (polymer/polymer blend)
(75% strain) (Figure4g).
Similar to the approach for an elastomeric composite
system for stretchable conductors, CPs/elastomer (e.g., sty-
rene-ethylene-butylene-styrene [SEBS], PDMS, or TPU) blend
systems have also been widely used. In particular, the similar
surface energies of the CPs and SEBS elastomer ensure a
nanoscale intimated morphology, which is crucial for ecient
charge transport in electronic devices. For examples, Xu etal.
reported intrinsically stretchable TFTs based on a DPP-based
CPs/SEBS active layer (Figure4h).[69] In these blended systems,
the increased polymer chain dynamics from nanoconfinement
significantly reduces the Young’s modulus of the CPs and
largely delays the crack formation under strain. Zheng et al.
also improved the chemical resistance of stretchable polymer
semiconductors by selectively crosslinking the backbones of the
elastomer (polybutadiene) and alkyl chains of DPP-based CPs to
make it feasible to photopattern stretchable polymer semicon-
ductors.[79] The fabricated TFTs showed a stable performance up
to 100% strain without aecting the charge carrier field-eect
mobility. In addition, stretchable polymer light emitting layers
with a nanoconfined light emitting polymer nanofiber network
in a polyurethane (PU) elastomer matrix showed stable photo-
luminescent characteristics on the skin during various defor-
mations (Figure4i).[64]
3. Stretchable Electrodes
Stretchable electrodes have been intensively investigated in
response to the rising demand for stretchable displays with new
form-factors such as foldable, rollable, and stretchable archi-
tectures. These displays must be able to operate reliably even
in unexpected deformations. Under such perspectives, many
researchers have developed stretchable electrodes to achieve
robust electrical interconnection between active devices such as
light-emitting devices and TFTs, since the active materials are
generally vulnerable to external stresses. To achieve mechanical
reliability and electrical conductivity simultaneously, various
methods with material-based and structural-based engineering
for stretchable electrodes have been developed, which can
maintain their interconnection under mechanical deformations
such as folding, stretching, and twisting circumstances.
In this regard, this section begins with a discussion of rep-
resentative approaches and fabrication technologies for stretch-
able electrodes, focusing on interconnection of active devices
on stretchable substrate, classifying them into 1) intrinsically
stretchable electrodes, 2) structurally designed stretchable elec-
trodes, and 3) vertical interconnection access (VIA) for highly
integrated stretchable system with advanced functionalities.
3.1. Intrinsically Stretchable Conductors
Stretchable electrodes with intrinsic elasticity are regarded as
one of the most promising candidates for realization of stretch-
able electronic systems. By maintaining conductive percolation
paths under mechanical deformation, the electrode with supe-
rior elasticity can endure localized strain from external stress,
and then successfully interconnect functional devices. As a rep-
resentative methods, elastomeric composites of an elastomer
and conductive fillers,[57,80–86] LMs,[87–93] and conductive poly-
mers[48,94–97] have been commonly used for intrinsically stretch-
able electrodes.
3.1.1. Conductive Nanomaterials Composites
As intrinsically stretchable electrodes, electrically conduc-
tive composites with combining an elastomer and conductive
fillers are commonly used. Depending on target stretchability
of desired applications, various types of elastomers such as
PDMS, PU, and fluorinated rubber can be used, and conduc-
tive fillers such as metal NWs,[80–83] metal particles,[57] and
CNT[84–86] can be mixed with the elastomer.
By using 1D metallic fillers, Lee et al. suggested AgNW-
PDMS nanocomposites as intrinsically stretchable elec-
trodes.[80] Since AgNW has 1D structure with high aspect ratio
as well as high electrical conductivity, the electrodes could
Adv. Mater. Technol. 2023, 2201067
2365709x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/admt.202201067 by Seoul National University, Wiley Online Library on [08/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com
© 2023 Wiley-VCH GmbH
2201067 (8 of 37)
www.advmattechnol.de
maintain electrical percolation paths by forming entangled
structure in an elastomer. To fabricate AgNW-embedded PDMS
structures, as-coated AgNWs on a carrier substrate were trans-
ferred to uncured PDMS, and then the intrinsically stretch-
able electrodes were formed by simultaneous embedding and
curing processes. The electrodes exhibited mechanical softness
due to the fact that the AgNW layer was formed just on the
surface of the PDMS substrate (Figure 5a), which has a neg-
ligible influence on elasticity of the substrate. When a tensile
strain of 20% was applied, the AgNW electrodes could absorb
the external strain and maintain their resistance under various
mechanical deformations such as bending and stretching.
There are several strategies for fabricating AgNW-based elec-
trodes, including direct patterning method such as inkjet
printing and bar coating as well as indirect fabrication tech-
niques such as shadow masking and dry transfer patterning.
Considering pros and cons of each fabrication methods, Yang
et al. reported facile and highly ecient fabrication methods
for robust AgNW–elastomer composite electrodes, combining a
vacuum filtration system with an improved 3D PDMS mask.[81]
The AgNW solution was successfully deposited on the target
region of membrane filter via the channels within the 3D mask.
As a result, clear edges could be achieved in both the AgNWs
deposited on the filter and transferred to the PDMS (Figure5b).
The resultant electrodes showed electrical conductivity of
3000 S cm1 at high deposition density of 10 g m2. By varying
the AgNW deposition density and the PDMS peel-o direction,
the stretchable electrodes could be stretched up to strain of
Figure 5. Stretchable composite electrodes. a) AgNW-embedded stretchable electrodes. Reproduced with permission.[80] Copyright 2020, Springer
Nature. b) AgNW–elastomer composite electrodes fabricated using vacuum filtration method. Reproduced with permission.[81] Copyright 2018, Royal
Society of Chemistry. c) Printable elastic electrodes with Ag flakes, fluorine rubber, and fluorine surfactant. Reproduced with permission.[57] Copy-
right 2015, Springer Nature. d) AgNW-embedded stretchable electrodes. Reproduced with permission.[84] Copyright 2019, American Chemical Society.
e) AgNW-elastomer composite electrodes fabricated using vacuum filtration method. Reproduced with permission.[85] Copyright 2022, American
Chemical Society.
Adv. Mater. Technol. 2023, 2201067
2365709x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/admt.202201067 by Seoul National University, Wiley Online Library on [08/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com
© 2023 Wiley-VCH GmbH
2201067 (9 of 37)
www.advmattechnol.de
80%, exhibiting mechanical robustness with the LEDs turning
on even under bending conditions.
Furthermore, since the composite mixed with elastomeric
matrix is generally in fluidic state before curing, various solu-
tion processes can be utilized to form stretchable electrodes
depending on target applications. In this regard, Matsuhisa
et al. reported a printable elastic electrode comprised of Ag
flakes, a fluorine rubber and a fluorine surfactant.[57] By modi-
fying the Ag flake surface from a water-based fluorine sur-
factant, which is main component for formation of conductive
networks, the electrical conductivity and mechanical stretch-
ability of the printed electrodes could be enhanced, showing
738 S cm1 at strain of 0% and 182 S cm1 at strain of 215%.
Although the electrical conductivity was relatively low, the
electrode with superior elasticity can be stretched over strain
of 200%, maintaining percolation paths of the composites by
adding the surfactant (Figure5c).
Metal nanomaterials have high electrical conductivity
because of their metallic nature but it also brings low mechan-
ical flexibility compared to carbon nanomaterials. By com-
bining those metal and carbon nanomaterials, they can have
synergetic eect on conductivity and stretchability. In this
regard, Ko etal. reported stretchable conductive interconnects
with superior electrical stability by forming composites that
consist of Ag particles with dierent sizes as 0D conductive
fillers, multi-walled CNTs (MWCNTs) as 1D conductive fillers
and silicone rubber matrix (Figure5d).[84] The micro-sized Ag
particles could enhance electrical conductivity of the compos-
ites, while the nano-sized Ag particles could allow dense con-
ductive network to be formed. Generally, the composite mixed
with only MWCNT networks has not enough electrical conduc-
tivity compared to counterparts with other metallic conductive
fillers. However, the MWCNTs used as auxiliary fillers could
enhance interactive attraction between Ag particles, enabling to
achieve high electrical conductivity and mechanical robustness
of the composites. The dierent-sized Ag particles in the sili-
cone matrix formed percolation paths and the MWCNTs were
positioned around the Ag particles. The subsequently achieved
stretchable electrodes exhibited high electrical conductivity of
6450 S cm1 with little change under 3000 tensile cycle tests
at strain of 50%. By utilizing aforementioned methods, Hong
et al. demonstrated intrinsically stretchable and printable lith-
ium-ion battery by using nanostructure-controlled multimodal
conductive fillers (Figure5e).[85] As the conductive fillers, var-
ious multimodal micro-particles such as Ag particles for the
anode and Ni particles for the cathode were used. As an auxil-
iary fillers, MWCNTs were also employed to form a networked
structure with conductive metal fillers. The composites with Ag
particles and Ni particles shows high electrical conductivity of
3912 and 2105 S cm1, respectively, which exhibited stable oper-
ation until tensile strain of 130%.
Conductive composites are widely used in apposite ways with
various conductive fillers and elastomeric substrates according
to the target application. However, the electrical conduc-
tivity remains low in compared to other metal film, since the
majority of components are nonconductive elastomer matrix.
In addition, the strain-sensitive electrical conductivity based on
percolation of conductive fillers is still limitation for cycling sta-
bility. Therefore, it is important to enhance mechanical stability
as well as electrical conductivity for implementation of practical
applications.
3.1.2. LMs
LM is a representatively used material for an intrinsically
stretchable electrode, because it can maintain high electrical
conductivity under strain by changing its shape from fluidic
property. Especially, since LM exists as liquid phase at room
temperature with ultralow Young’s modulus, an LM elec-
trode is capable of easily deformed to any direction with neg-
ligible resistance change unlike other solid-phase stretchable
electrodes.
Taking advantage of fluidic property, Yang et al. used a
eutectic alloy of LM (Galinstan) that consists of Ga, In, and tin
(Sn) for intrinsically stretchable electrodes.[87] The LM could
flow in a liquid form, congregating together with high surface
tension (Figure 6a). Since the electrical conductivity can be
maintained due to self-healing property as long as the inter-
connection is not completely damaged, the stretchability of
LM electrodes can be directly aected by the material proper-
ties of an elastomeric substrate such as Young’s modulus, yield
strength, and maximum yield point strain. To achieve high
stretchability over 300%, the Ecoflex silicone rubber was used
as an elastomeric substrate, and then the LM was injected into
empty channel in the rubber to form electrical interconnection.
The subsequently achieved LM electrodes showed superior
stretchability over 300% of strain.
Similarly, Bartlett et al. also reported LM-embedded elasto-
mers. Mixing the LM eutectic alloy that consists of Ga and In
with Ecoflex or PU elastomer, the intrinsically stretchable elec-
trodes were fabricated (Figure6b).[88] Various types of solution
fabrication processes including dispensing, screen printing,
stencil printing, and spray coating can be utilized for patterning
of the LM-elastomer composites. Especially, the LM electrodes
patterned by stencil printing exhibited superior stretchability
from 0% to 500% strain. To further enhance mechanical reli-
ability, Pan et al. suggested a LM-elastomer composite with
various sizes of LM droplets in an elastomeric substrate
(Figure6c).[89] The eect of LM droplet sizes to stiness of the
composites was investigated. The LM-elastomer composites
with small fillers could maintain mechanical stretchability, but
counterpart composites with large fillers showed a degradation
of stretchability in the strain at break. Therefore, the sizes of
LM droplets must be carefully considered depending on target
stretchability.
As another characteristic of LM electrodes, a thin oxide layer
can be naturally and rapidly formed around the LM electrodes,
enabling various architectures to be easily designed with main-
taining mechanical stability. However, there are challenges in
high resolution patterning due to its rapid oxidation with high
surface tension in air environment. To overcome this issue
and exhibit feasibility of the LM electrodes for stretchable
electronic systems with various designs, Silva et al. reported
high-resolution stretchable circuits with finely patterned LM
electrodes by controlling the oxide layer of LM (Figure6d).[90]
Polyvinyl alcohol (PVA) was coated on an elastomeric substrate,
which enables inkjet printing of conductive materials on an
Adv. Mater. Technol. 2023, 2201067
2365709x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/admt.202201067 by Seoul National University, Wiley Online Library on [08/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com
© 2023 Wiley-VCH GmbH
2201067 (10 of 37)
www.advmattechnol.de
elastomeric substrate, followed by printing of Ag nanoparticles
(AgNPs). By removing oxide layer of LM through dissolving in
acetic acid solution, the LM could be selectively patterned on
the printed AgNP traces due to its dewetting property from the
PVA layer. Taking advantages of the self-patterning characteris-
tics, complicated patterns and circuits with dierent fine lines
could be achieved, exhibiting outstanding electrical conductivity
and stretchability over 100% of strain.
Although LM electrode can exhibit superior elasticity with
high electrical conductivity, there are still critical issues to use it
together with other conductive metal, since Ga easily penetrates
other metals along to grain boundaries, resulting in irreversible
performance degradation such as swelling and cracking. There-
fore, specific integration methods including pattering, deposition,
and encapsulation technologies must be considered together.
3.1.3. Conductive Polymers
Conductive polymers have advantages in direct processability
and tunability of the molecular structures to determine elec-
trical and mechanical properties. Since they are capable of
being soluble in proper solvents, various solution processes
such as inkjet printing, spray coating, and screen printing can
be utilized. Among a lot of conductive polymers, PEDOT:PSS
has been intensively utilized as transparent stretchable elec-
trode for stretchable electroluminescent devices due to its high
electrical conductivity, transparency, and easy work function
modification compared to other conductive polymers.
Oh et al. reported modification of nanostructure and vis-
coelastic property of PEDOT:PSS by using plasticizer (Triton
X-100).[94] By optimizing weight fraction of the surfactant,
electrical conductivity and mechanical stretchability could
be enhanced simultaneously. Due to the viscoelasticity from
the addition of Triton X-100, the conductive polymer could be
molded into micro-patterns which enable implementation of
high resolution OLEDs array (Figure 7a). PEDOT:PSS elec-
trodes are generally incorporated with such plasticizers to
enhance its stretchability, but the electrical conductivity is still
low to be used for realization of high performance stretch-
able devices. To further improve stretchability and electrical
conductivity, Wang et al. reported the highly stretchable,
transparent and conductive PEDOT:PSS electrodes, which
were incorporated with ionic additives-based stretchability
and electrical conductivity (STEC) enhancers (Figure 7b).[48]
The STEC enhancers with ionic additives enabled forma-
tion of soft domains, better connectivity between PEDOT-
rich domains, and higher crystallinity in PEDOT regions,
resulting in implementation of highly conductive and stretch-
able PEDOT:PSS electrodes which exhibited 4100 S cm1
under 100% of strain.
Similarly, Park et al. suggested transfer-printed conduc-
tive polymer electrodes for highly customizable PLEDs.[95] To
achieve better wettability on an elastomeric substrate, surface
treatments such as ultraviolet (UV) light exposure and oxygen
plasma treatment have been widely used, but these treatments
could degrade performance of organic layers. To overcome
this issue, they suggested dry transfer printing methods for
Figure 6. LM stretchable electrodes. a) A eutectic alloy of LM that consists of Ga, In, and Sn for intrinsically stretchable electrodes. Reproduced with
permission.[87] Copyright 2018, American Chemical Society. b) LM microdroplet-embedded stretchable electrodes. Reproduced with permission.[88]
Copyright 2016, Wiley-VCH. c) LM–elastomer composites with various sizes of LM droplets. Reproduced with permission.[89] Copyright 2019, Wiley-
VCH. d) Finely patterned LM stretchable electrodes for high-resolution applications. Reproduced with permission.[90] Copyright 2020, Wiley-VCH.
Adv. Mater. Technol. 2023, 2201067
2365709x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/admt.202201067 by Seoul National University, Wiley Online Library on [08/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com
© 2023 Wiley-VCH GmbH
2201067 (11 of 37)
www.advmattechnol.de
PEDOT:PSS patterning to minimize degradation of organic
layers during whole fabrication processes. The PEDOT:PSS
electrodes incorporated with D-sorbitol, which was used to
enhance bonding force in a lamination processes, were suc-
cessfully transferred to a PDMS substrate. The transferred elec-
trodes exhibited outstanding transparency compared to conven-
tional ITO (Figure 7c). As a proof of concept, the matrix dis-
play were demonstrated which maintains its brightness under
bending state when the PEDOT:PSS electrodes were used as
anodes of PLEDs.
Despite its transparency and intrinsic elasticity, since
PEDOT:PSS has relatively low electrical conductivity com-
pared to metal film, it is dicult to use it alone as an stretch-
able electrode. Therefore, multidisciplinary research should be
conducted to improve electrical and mechanical performance
through material synthesis or hybrid with other electrodes.
3.2. Structurally Designed Stretchable Interconnects
Structurally designed stretchable electrodes with various geo-
metrical patterns are one of the most promising candidates for
interconnection component of high-performance stretchable
display, taking advantages originated from well-established fab-
rication processes and material properties of metal films with
high electrical conductivity.[98] However, since the metal films
such as Cu, Au, and Ag have much lower tensile limit than
intrinsically stretchable electrodes, specific structural designs to
dissipate localized strain should be employed. Therefore, many
researchers have developed representative structural designs
such as serpentine[99–102] and wrinkled electrodes,[98,103–105]
which can maintain electrical conductivity by changing their
geometrical shapes.
3.2.1. Serpentine Electrodes
Serpentine electrodes are commonly used as metallic stretch-
able electrodes due to their superior electrical conductivity. The
metal electrodes are capable of being easily patterned by well-
established conventional photolithographic processes. Due to
unique structural designs with horseshoe shapes, the electrodes
could endure localized strain by changing their geometrical
shapes under various mechanical deformations (Figure 8a).[99]
Considering this characteristic, Biswas etal. reported stretch-
able active matrix display with multilayered serpentine elec-
trodes.[100] By using photolithography process, 10 µm thick
Cu electrode was deposited on carrier substrate where sacrifi-
cial layers of poly(methyl-methacrylate) (PMMA) and PI were
coated, and then insulation layer composed of photopatternable
PI was used to secure separation of row and column electrodes
(Figure8b). Even though there were thick parts of coated films,
the electrodes could be stretched due to its structural design.
To demonstrate their stretchability, commercially available sur-
face mount LEDs were integrated with the electrodes, followed
by encapsulation process with Ecoflex elastomer with very low
Young’s modulus to achieve softness. The subsequently fabri-
cated stretchable active matrix display exhibited no significant
performance changes under stretching at 220% of strain.
By further developing the serpentine structure designs, Li
et al. suggested 2D fractal-inspired shapes for advanced inter-
connect configurations.[101] The unique shapes could render
metal electrodes enhanced stretchability compared to conven-
tional serpentine electrodes, but the stretchability is still as low
as 11% when the electrodes were just stretched along to in-plane
direction. Therefore, two-stage solid encapsulation methods
were further suggested for formation of 3D configurations,
resulting in four times improved stretchability up to 46% strain
Figure 7. Conductive polymer-based stretchable electrodes. a) Micropatterned PEDOT:PSS electrodes blended with Triton X-100 to control viscoe-
lasticity. Reproduced with permission.[94] Copyright 2016, Wiley-VCH. b) PEDOT:PSS-based stretchable interconnects with ionic additives and STEC
enhancers. Reproduced with permission.[48] Copyright 2017, American Association for the Advancement of Science. c) PEDOT:PSS electrodes incorpo-
rated with D-sorbitol to enhance bonding force during lamination processes. Reproduced with permission.[95] Copyright 2019, Wiley-VCH.
Adv. Mater. Technol. 2023, 2201067
2365709x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/admt.202201067 by Seoul National University, Wiley Online Library on [08/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com
© 2023 Wiley-VCH GmbH
2201067 (12 of 37)
www.advmattechnol.de
(Figure8c). By using these configurations, stretchable LED sys-
tems could be successfully achieved, maintaining electrical per-
formance even under 40% of biaxial strain.
3.2.2. Wrinkled Electrodes
Wrinkled electrodes can be directly formed on elastomeric
substrates using not only vacuum evaporation but also various
solution processes such as inkjet printing, screen printing,
spray coating, and electroless deposition. Among various
methods, inkjet printing is commonly used for patterning of
the metal films due to its facile customizability and low temper-
ature processability. Cho etal. reported strain-tolerant wrinkled
electrodes for stretchable electrodes.[103] The wrinkled structure
could be introduced by releasing a pre-stretched elastomeric
substrate where conductive inks were printed by inkjet printing
method. Especially, they pointed out that the stretchability was
dominantly aected by the regularity of wrinkles since mechan-
ical strain could be concentrated at the thicker regions of the
electrode. By optimizing inkjet printing processes for imple-
mentation of thin and flat metal films, well-defined wrinkles
Figure 8. Structurally designed serpentine and wrinkled stretchable electrodes. a) Horseshoe-shaped serpentine structural designs to endure localized
strain. Reproduced with permission.[99] Copyright 2014, Wiley-VCH. b) Serpentine stretchable electrodes fabricated using well-established conventional
photolithographic processes. Reproduced with permission.[100] Copyright 2019, Springer Nature. c) Fractal-inspired serpentine electrodes for advanced
interconnect configurations. Reproduced with permission.[101] Copyright 2019, Wiley-VCH. d) Strain-tolerant wrinkled electrodes. Reproduced with
permission.[103] Copyright 2019, Taylor & Francis. e) Wrinkled electrodes with integrated LEDs fabricated by customizable inkjet printing methods.
Reproduced with permission.[104] Copyright 2017, Springer Nature. f) Corrugated SWCNT diusion barrier with wrinkled electrodes. Reproduced with
permission.[105] Copyright 2018, Wiley-VCH.
Adv. Mater. Technol. 2023, 2201067
2365709x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/admt.202201067 by Seoul National University, Wiley Online Library on [08/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com
© 2023 Wiley-VCH GmbH
2201067 (13 of 37)
www.advmattechnol.de
on the films could be obtained (Figure8d), and showed a little
resistance changes of 10% even under 10 000 stretching cycles
at 25% tensile strain. Similarly, Byun et al. fabricated various
types of stretchable display by using customizable inkjet
printing methods with integrating LEDs on wrinkled elec-
trodes.[104] By forming gradual wrinkles on metal films with
embedded strain modulators, the mounted LEDs could main-
tain their electrical performance on human skin (Figure8e).
To further improve mechanical reliability, Oh etal. suggested
a corrugated SWCNT diusion barrier for the wrinkled elec-
trodes.[105] LM was used as contact materials, which could dis-
sipate concentrated mechanical strain at the boundary between
LEDs and wrinkled electrodes. However, the LM could react
with the wrinkled electrodes by penetrating grain boundaries in
metals. Considering this issue, corrugated SWCNT was intro-
duced to diusion barrier for wrinkled electrodes. Based on
the corrugated and wrinkled configurations, stretchable display
with wrinkled electrodes was demonstrated, which showed no
noticeable change in the luminance of the LEDs under 20% of
biaxial strain (Figure8f).
As described in previous examples, stretchable electrodes
with serpentine and wrinkled structure could be variously
designed depending on target stretchability of desired applica-
tions, and exhibited outstanding performance with high elec-
trical conductivity under various mechanical deformations.
However, the fabrication processes for serpentine electrodes
such as photolithography and etching are still challenging
for direct fabrication onto elastomeric substrates, and pre-
stretching process for wrinkled structures is critical bottleneck
for mass production of stretchable display. Therefore, it is
important to further establish suitable fabrication technologies
considering compatibility, processability, customizability, and
feasibility.
3.3. VIAs with Stretchable Electrodes
To implement a stretchable matrix display integrated with mul-
tifunctional layers, formation of VIAs in stretchable substrate
has been vigorously investigated in multidisciplinary research
fields. By interconnecting various active layers in stretchable
systems with high integration density, versatile operations
can be achieved. To form VIAs in stretchable substrate, mate-
rials and fabrication technologies should be carefully selected
since conductive paths along to vertical direction must be reli-
ably maintained when the systems are stretched along to hor-
izontal direction. In addition, the VIAs are generally filled in
perforated regions in an elastomeric substrate, which can lead
electrical failure and mechanical tear at the regions. There-
fore, many research groups have considered various methods
to make VIAs strain-insensitive without degrading mechanical
reliability and stretchability of stretchable substrates.
To form VIAs with serpentine electrodes, Huang et al.
reported a framework for 3D integrated stretchable systems
with VIAs, which were formed by laser ablation and filling con-
ductive materials in predefined region.[106] Elastomer matrix
were selectively perforated by laser ablation with careful opti-
mizations of fabrication parameters such as laser wavelength,
pulse, and absorption rate. The geometries of VIAs such as
depth and width were controlled to secure electrical contacts
between VIA and other layer. By increasing diameter from
300 to 600 µm up to the fourth layer, enough contact in through
VIA could be ensured by filling conductive solder pastes in the
region (Figure 9a). Although the VIA was nonstretchable due to
rigid nature of filling material, it was stably interconnected by
serpentine electrodes with island-bridge configurations. Based
on this approach, highly conductive and strain-insensitive
VIAs could be achieved, and the VIAs exhibited ultralow
Figure 9. VIAs with stretchable electrodes. a) 3D integrated stretchable systems with VIAs formed by laser ablation and filling conductive materials with
serpentine electrodes. Reproduced with permission.[106] Copyright 2018, Springer Nature. b) Modulus-gradient stretchable VIAs fabricated by printing-
based methods with wrinkled electrodes. Reproduced with permission.[107] Copyright 2017, Wiley-VCH. c) Liquid alloy stretchable VIAs fabricated by
spray coating methods with LM stretchable electrodes. Reproduced with permission.[108] Copyright 2021, Wiley-VCH.
Adv. Mater. Technol. 2023, 2201067
2365709x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/admt.202201067 by Seoul National University, Wiley Online Library on [08/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com
© 2023 Wiley-VCH GmbH
2201067 (14 of 37)
www.advmattechnol.de
resistance about 68.7 m, which enables implementation of high-
performance stretchable systems with increased integration
density.
In case of VIA with wrinkled electrodes, Byun et al. sug-
gested double-side soft electronic platform with core–shell
VIA.[107] By using printing-based method, the conductive VIA
was patterned on an elastomeric substrate. To resolve discon-
tinuity in elastic modulus between the VIA and substrate,
which could lead unexpected delamination, geometrical design
and compositions of the VIA were carefully optimized. The
conductive composites that consist of Ni particles, PDMS,
and silicone resin were printed on desired position, and then
magnetic alignment of the composites was conducted to form
gradual stiness in a single VIA (Figure9b). Due to the focused
magnetic alignment, Ni particles were converged into center
of VIA and aligned along to vertical direction, simultaneously.
The subsequently stiness-gradient shell, being relatively high
Young’s modulus than the substrate, could stably protect the
conductive vertical paths. Based on this approach, double-side
stretchable platform with wrinkled electrodes was successfully
demonstrated, which showed high-speed operation of stretch-
able computing circuits.
To implement LM-based VIA, Jiang etal. reported LM alloy
VIAs for fabricating multilayered stretchable systems.[108] To
generate microholes in an elastomeric substrate, UV laser for
ablation was used to implement vertical profile with microstruc-
tures, and then the liquid alloy was deposited in the holes by
spray coating. Taking advantages in unique properties such as
liquid phase and rapid oxidation in air environment, the liquid
alloy could be coated on both sides of holes (Figure9c). After
selectively pinning at both sides, the oxide layer surrounding
the coated liquid alloy was rapidly formed, which enable that
the liquid alloy could maintain its shape against high surface
tension and gravity. The subsequently formed VIA exhibited a
resistance variation (ΔR/R0<300%) under stretching test with
tensile strain >90%, which is average strain of elongation at
break for elastomer substrate.
As aforementioned examples of VIA for stretchable system,
there are various fabrication methods and material candidates
to achieve robust interlayer interconnection and high integra-
tion in the stretchable system with multilayered structure. To
further improve integration density with maintaining electrical
performance and mechanical stability simultaneously, apposite
strategies must be selected depending on target applications
with careful consideration of the pros and cons in each method.
4. Stretchable TFTs
Stretchable TFT is a fundamental component in the backplane
of stretchable display, driving the electroluminescent devices
as well as performing signal switching and amplification. To
implement advanced stretchable systems, there are signifi-
cant eorts to fabricate mechanically reliable stretchable TFTs,
which can maintain their performances in various mechanical
deformations. Along with stretchable electrodes that intro-
duced in the previous chapter, the development of stretch-
able semiconducting active materials is critical for stretchable
TFTs. Therefore, it is important to develop apposite strategies
for implementation of stretchable TFTs depending on active
channel materials. In this regard, we discuss the various meth-
odology of forming stretchable TFTs in terms of active channel
materials, classifying into 1) inorganic, 2) CNT, and 3) organic
semiconductors.
4.1. Inorganic TFTs
Despite the inherent brittleness of inorganic active channel
materials including conventional silicon-based materials and
oxide semiconducting materials, there are various attempts
to implement stretchable inorganic TFTs for practical appli-
cations. Kim et al. suggested wrinkled structured TFTs for
silicon-based CMOS integrated circuits.[17] Although the inor-
ganic materials are generally vulnerable to external stress, the
wrinkled structures can render the fragile materials stretchable
by minimizing the deformation energy. Since inorganic TFTs
could be hardly fabricated onto the elastomeric substrate due
to the high temperature process, the TFTs were transferred
from the carrier substrate to prestretched PDMS substrate, and
then released them to initial state to achieve desired stretch-
ability. As proof of concept, stretchable CMOS integrated cir-
cuits were demonstrated, which could maintain their electrical
performances under deformations (Figure 10a). Although the
wrinkled inorganic TFTs has excellent electrical properties due
to well-defined crystallinity of inorganic semiconductors, they
require specialized manufacturing procedures such as pre-
stretching and transferring. In addition, cracks are also likely
to occur when external strain applied to the device exceed limit
of prestrain, which can cause irreversible damage of electrical
behavior. Moreover, since TFTs consist of multiple layers with
electrodes, dielectric, and active channels, it is dicult to form
regular wrinkles on whole layers due to variation of their thick-
ness and Young’s moduli.
A strain-modulating structures such as an island-bridge
structure can be an ecient way to facilitate fabrication pro-
cess and to modulate strain distribution.[109–112] Since the inor-
ganic semiconducting materials have higher Young’s moduli
compared to elastomeric substrate, they can be located on the
strain-free region where the rigid parts are formed in the soft
substrate. When TFTs are fabricated on the strain-free areas
with stretchable interconnects, the deformation of TFTs can be
minimized. From this perspectives, Park et al. demonstrated
an array of stretchable oxide TFTs by transferring ZnO TFTs
onto the strain-modulated substrate with rigid island struc-
ture (Figure10b).[109] The stretchable interconnects with wrin-
kled structures could endure bidirectional strain, reducing the
mechanical stress on brittle oxide TFTs formed on strain free
regions.
By further developing material and structural designs, Park
et al. used LM for stretchable electrodes and fabricated TFTs
at the cross-point rigid island (Figure 10c).[110] The LM-based
stretchable electrodes can improve maximum strain limita-
tions due to its fluidic properties with maintaining electrical
conductivity under various deformations. Moreover, LM can
be easily patterned in submicrometer feature size.[113] Also, to
improve scalability of stretchable TFTs, Cantarella et al. sug-
gested structurally engineered substrate where pillar structures
Adv. Mater. Technol. 2023, 2201067
2365709x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/admt.202201067 by Seoul National University, Wiley Online Library on [08/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com
© 2023 Wiley-VCH GmbH
2201067 (15 of 37)
www.advmattechnol.de
were formed to obtain strain-free regions (Figure10d).[111] Since
the pillar structures could be freely designed according to the
geometries of designed TFTs, it could be applied to large-scale
logic circuits which consists of various types of TFTs.
Likewise, most of researches about stretchable inorganic
TFTs have generally focused on structural designs to achieve
mechanical stretchability and robustness. However, structural
improvement alone cannot overcome the intrinsic brittleness
fundamentally. In addition, conventional fabrication methods
such as photolithography and etching are not fully compatible
with elastomeric substrates, resulting in further increasing the
diculty of the whole processes. For these reasons, additional
integration techniques should be developed for realization of
more advanced stretchable systems.
4.2. CNT TFTs
CNT is one of the promising materials for implementation of
stretchable TFTs due to its superior electrical and mechanical
properties.[114–118] The 1D nanostructured CNTs can maintain
their electrical percolation paths with formation of entangled
networks.
In this regard, Chae et al. demonstrated stretchable CNT
TFTs by combining semiconducting CNT networks, graphene
electrodes, and wrinkled Al2O3 dielectric layer.[117] In contrast to
conventional inorganic materials, since the CNT networks and
graphene layers are extremely thin, well-defined wrinkles could
be easily formed. By combining CNTs with corrugated structure
design of dielectric layer, the mechanical stability could be fur-
ther improved without compromising electrical characteristics
up to 20% strain in channel width direction, which showed
on/o ratio of 105 and high mobility of 40 cm2 V1 s1. This
approach not only showed the structural design applicability
of carbon-based materials but also exhibited their stretchability
and feasibility for intrinsically stretchable TFTs. In addition,
CNTs can be used for both electrodes and active channel mate-
rials simultaneously depending on their metallic and semicon-
ducting properties, respectively. In this regard, Cai et al. dem-
onstrated fully printed stretchable TFTs and integrated logic
circuits by using CNTs as both electrode and active channel
materials (Figure 11a).[119] To implement fully stretchable TFTs,
they fabricated stretchable dielectric composite that consist
of barium titanate (BaTiO3) NPs and PDMS matrix, which
exhibited high dielectric constant and stretchability. Due to
the intrinsic stretchability of electrodes, dielectric, and active
Figure 10. Stretchable inorganic transistors. a) Wrinkled structured CMOS circuit based on silicon active channel and its output characteristics with
external strain Reproduced with permission.[17] Copyright 2008, American Association for the Advancement of Science. b) Image of island structure
used for stretchable oxide transistor array, with 5% of x- and y-axis strain. Interconnections are formed with a wrinkle to withstand the external strain.
Reproduced with permission.[109] Copyright 2010, Wiley-VCH. c) Image of island-structured oxide transistors with 40% external strain. Interconnections
are LM-based electrodes. Reproduced with permission.[110] Copyright 2018, The Japan Society of Applied Physics. d) Schematic of large pillar islands
used to fabricate a transistor array or logic circuits. Reproduced with permission.[111] Copyright 2018, Wiley-VCH.
Adv. Mater. Technol. 2023, 2201067
2365709x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/admt.202201067 by Seoul National University, Wiley Online Library on [08/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com
© 2023 Wiley-VCH GmbH
2201067 (16 of 37)
www.advmattechnol.de
channels, the TFTs showed stretchability beyond 50% of strain
along direction of length or channel width, and exhibited no
significant degradation of electrical performance during tensile
stretching cycles.
To fabricate high performance stretchable CNT TFTs with
high mobility, low voltage, and high integration density,
Huang etal. combined the advantages of solution process and
conventional photolithography methods (Figure11b).[120] They
used poly(urea-urethane) (PUU) elastomer as stretchable die-
lectric material because it has a higher permittivity compared
to other elastomeric dielectrics. However, the PUU is easily
damaged from photolithographic processes such as plasma
etching. To overcome this limitation, removal-transfer-
photolithography method was proposed. As a sacrificial layer,
indium gallium zinc oxide (IGZO) was used to enhance
wettability of PDMS substrate and PUU dielectric, and
etching damage of the PUU dielectric was minimized with the
appropriate use of sacrificial layers. Since CNT networks have
superior carrier mobility and the charge transportation was
significantly improved by controlling dielectric capability, the
stretchable CNT TFTs showed high mobility (221 cm2 V1 s1)
and maintained their characteristics after 2000 stretching
cycles with 50% strain.
Another advantage of CNT to expand the scope of device
applications is transparency, which can increase aperture
ratio of display pixels or be used for fully transparent wear-
able devices. Liang etal. fabricated the intrinsically stretchable
and transparent CNT TFTs (Figure11c).[121] AgNWs and PU-co-
polyethylene glycol were used for electrodes and elastomeric
dielectric for intrinsically stretchable and transparent devices.
The fabricated TFT could be stretched up to 50% strain with
maintaining its electrical performance and exhibited optical
transmittance over 90% in the 450–1100 nm wavelength range,
driving OLED in the full brightness range. These results
implied the feasibility of a CNT-based backplane for display
devices with high transparency.
Furthermore, CNT TFTs could be integrated with stretch-
able sensors for various wearable devices. Hong et al. fabri-
cated a skin-attachable and intrinsically stretchable CNT TFT
array with highly sensitive temperature sensors.[122] By using
thermos-responsive film, CNT TFT arrays with a thermo-
chromic property in the range of human body temperature
were implemented (Figure 11d). The fabricated devices were
easily attached to the human skin with stable operation under
skin movement such as wrist bending. As a practical healthcare
application, thermochromic display devices with CNT TFTs
were demonstrated with showing a possibility of advanced
wearable display application.
With the aforementioned unique characteristics, CNTs
showed the great possibility to be applied in stretchable driving
devices. However, in the chemical manufacturing process of
CNTs, both metallic and semiconducting CNTs are grown at
once. Therefore, additional purifying processes are needed to
improve the semiconducting property of the active channels.
Figure 11. Stretchable CNT transistors. a) Schematic and image of intrinsically stretchable CNT-based transistors fabricated using inkjet printing
process. Reproduced with permission.[119] Copyright 2016, American Chemical Society. b) Stretchable CNT transistors with high resolution achieved by
photolithography. Reproduced with permission.[120] Copyright 2020, Royal Society of Chemistry. c) OLED operation with various bias and strain condi-
tions. Interconnections are LM-based electrodes. Reproduced with permission.[121] Copyright 2015, Springer Nature. d) Schematic of a thermochromic
wearable display fabricated using CNT transistors. Reproduced with permission.[122] Copyright 2019, Wiley-VCH.
Adv. Mater. Technol. 2023, 2201067
2365709x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/admt.202201067 by Seoul National University, Wiley Online Library on [08/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com
© 2023 Wiley-VCH GmbH
2201067 (17 of 37)
www.advmattechnol.de
Because the remaining metallic CNTs form leakage paths in
the channel even after purification, the o-current level can be
increased. In addition, the random network structure of CNT
channels reduce the uniformity of TFT array. Therefore, devel-
opment of purification methods with high yield and studies of
advanced passivation and percolation structure are required to
achieve CNT TFT array with high stability and uniformity.
4.3. Organic TFTs
Organic semiconductors are also one of the promising candi-
dates for active layers of deformable TFTs based on not only
easy processability including vacuum and solution processes,
but also easy controllability of mechanical and electrical prop-
erties through molecular engineering. Similar to the inorganic
TFTs, the island-bridge structures were implemented in the
early stages of researches about stretchable organic TFTs due to
mechanical brittleness of small molecule organic semiconduc-
tors.[61,123] Sekitani etal. demonstrated a stretchable organic TFT
array utilizing PI as rigid islands (Figure 12a).[123] The brittle
channel region was located on strain-free region with PI rigid
islands to minimize the mechanical stress when the device was
stretched. As a result, the fabricated TFT array with stretchable
interconnects could be uniaxially and biaxially stretched up to
70% without any mechanical or electrical failures. However,
Figure 12. Stretchable organic transistors. a) Schematic of stretchable organic transistor array utilizing PI as rigid islands and images of fabricated
devices under stretched states. Reproduced with permission.[123] Copyright 2008, American Association for the Advancement of Science. b) Schematic
of polymer dynamics of block copolymer with amorphous and crystal chains and a self-healing process with solvent vapor treatment. Reproduced with
permission.[125] Copyright 2016, Springer Nature. c) Changing the crystallinity of conjugated polymers by addition of a conjugated carbon nanoring.
Reproduced with permission.[132] Copyright 2019, Wiley-VCH. d) Schematic of solution-shearing process for a composite of elastomer and unidirec-
tional-ordered conjugated polymer induced by nanoconfinement eect. Reproduced with permission.[134] Copyright 2019, Springer Nature. e) Schematic
of stretchable organic transistor array with composite-based stretchable semiconductor and elasti layer for redistribution of applied strain. Repro-
duced with permission.[135] Copyright 2021, Springer Nature.
Adv. Mater. Technol. 2023, 2201067
2365709x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/admt.202201067 by Seoul National University, Wiley Online Library on [08/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com
© 2023 Wiley-VCH GmbH
2201067 (18 of 37)
www.advmattechnol.de
as mentioned in the section of inorganic and CNT transistors,
this approach utilizing rigid islands has still limitations about
process complexity due to additional processes for fabrication
of rigid regions in the elastomeric substrate and concentrated
mechanical stress at the interface.
To overcome aforementioned limitations, molecular engi-
neering methods for the implementation of intrinsically
stretchable organic semiconducting materials have been devel-
oped. There are three main strategies to develop intrinsically
stretchable organic semiconductors: copolymerization or side
chain engineering to control crystallinity,[124–130] blending addi-
tives to induce phase separation or change in crystallinity,[131,132]
and making composite of organic semiconductor and elas-
tomer.[66,69,133–135] In the first case, Oh etal. demonstrated intrin-
sically stretchable and healable block copolymer-based organic
TFTs with conjugated repeating units and non-conjugated moi-
eties as hydrogen bonding units.[125] When strain was applied
to the polymer films, energy dissipation occurred through
the breakage of hydrogen bonds between amorphous chains
(Figure12b). As a consequence of this energy dissipation, the
polymer semiconductor could retain high charge transport abil-
ities in stretched states. Furthermore, the mechanically dam-
aged polymer film could be healed through solvent vapor treat-
ment and thermal annealing, recovering its morphology and
field-eect mobility to those of an undamaged film. Stretchable
organic TFTs fabricated with CNT electrodes and PDMS dielec-
tric layer could operate stably under 100% strain and even after
the 500 stretching cycles at 25% strain.
In the second case, as a strategy of additive to control the
crystallinity of polymers, Mun et al. blended a conjugated
carbon nanoring, cycloparaphenylenes (CPP), to the conjugated
polymer semiconductor to tune the dynamic behaviors of the
polymer.[132] As CPP was added to the polymer, long-range crys-
talline order was reduced, thereby enhancing polymer dynamic
motion (Figure 12c). In addition, CPP improved the device
performance by lowering the contact resistance and increasing
charge transportation. Subsequently fabricated stretchable TFTs
with CNT electrodes and PDMS dielectric layer could operate
without any dramatic degradation under 100% strain and
1000 stretching cycles with 25% strain.
Although the two main strategies mentioned above enable
the realization of intrinsically stretchable organic TFTs, there
are some limitations on fine-tuning of stretchability and elec-
trical characteristics of organic semiconductors simultaneously
due to constraints of molecular engineering. To overcome those
limitations, researchers utilize conjugated polymer and elas-
tomer composites to give stretchability to the semiconductors.
Xu etal. reported a composite-based stretchable organic semi-
conductor that conjugated polymers in composite was unidirec-
tionally ordered by a solution-shearing process (Figure12d).[134]
A nanofiber structure of conjugated polymer in composite was
achieved from a nanoconfinement eect induced by a phase
separation of elastomer and conjugated polymer. In addition,
polymer chains in the composite were aligned and elongated
by the solution-shearing process with the intensive unidirec-
tional flow generated from microtrenches of the coating blade.
Due to the multi-scale ordering and alignment of conjugated
polymers in stretchable composite, charge carrier mobility
and stretchability of stretchable semiconductors were greatly
enhanced. As a result, fabricated stretchable organic TFTs with
CNT electrodes and SEBS dielectric layer can operate without
any failures under 100% strain and 1000 stretching cycles with
50% strain. Furthermore, Wang et al. introduced patterned
elastomer layers with tunable stiness to protect composite-
based organic semiconductors from the applied strain.[135] By
locally tuning the cross-linking density of the elastomer, sti
parts of the substrate were utilized as strain-relief regions sim-
ilar to rigid island structures (Figure 12e). By combining two
approaches of strain-relief structures and composite of elas-
tomer/conjugated polymer, the fabricated intrinsically stretch-
able organic TFTs showed strain-insensitivity with the change
of performance less than 5% under 100% strain.
As described above, organic semiconductors have signifi-
cant advantages of controlling their mechanical and electrical
properties by engineering molecular structures or blending
several organic materials. These organic TFTs can achieve
high stretchability without any complicated approaches such
as island-bridge structures or kirigami patterns. On the other
hand, organic semiconductors have disadvantages of vulnera-
bility to chemicals, UV irradiation, or plasma treatments. How-
ever, these limitations are also becoming overcome by selec-
tive molecular crosslinking of organic materials. Recent work
reported high-density intrinsically-stretchable transistors and
circuits with all-polymer-materials by direct optical microlithog-
raphy patterning with channel length of 2 µm and a density of
42 000 transistors cm2.[136] The electrical properties of polymer
semiconductors are lower than inorganic- and CNT-based coun-
terparts, but the intrinsically-stretchable polymer transistor
array can be a potential candidate for wearable and electronic
skin devices.
5. Stretchable Light-Emitting Devices
Stretchable light-emitting devices stably maintain light-emis-
sion under various deformations, providing new design form-
factors for deformable electronics, wearable electronics, and soft
robotics. There are two main strategies to demonstrate stretch-
able electroluminescent devices: adapting island-bridge config-
uration to the rigid and flexible electroluminescent devices and
developing intrinsically stretchable electroluminescent devices.
Based on island-bridge configurations, various light-emitting
devices are integrated with stretchable interconnects that can
accommodate external stresses. When external strain is applied,
the strain is concentrated into stretchable interconnects, while
the active devices that are positioned in the strain-free region
can be eectively protected, exhibiting improved mechanical
stability. In the case of intrinsically stretchable electrolumi-
nescent devices, approaches are similar to the development
of intrinsically stretchable TFTs, such as blending additives
to induce phase separation or making a composite of light-
emitting materials with elastomers or low glass transition
temperature polymers to give stretchability. Each strategy for
structure and material engineering to achieve desired stretch-
ability can be determined based on the target light-emitting
materials. In this section, the representative strategies for the
development of stretchable light-emitting devices will be pre-
sented with the various light-emitting materials and devices,
Adv. Mater. Technol. 2023, 2201067
2365709x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/admt.202201067 by Seoul National University, Wiley Online Library on [08/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com
© 2023 Wiley-VCH GmbH
2201067 (19 of 37)
www.advmattechnol.de
including 1) inorganic, 2) polymer (organic), and 3) perovskite
light-emitting materials and LEDs, and 4) ACEL.
5.1. Inorganic LEDs
Inorganic LEDs (ILEDs) are representative electroluminescent
devices that consist of inorganic compounds of ‘III-V’ or ‘II-VI’
materials. Combinations of Ga, In, N, and P such as GaN[137]
and GaAs[138] work as a p–n junction with radiative recombina-
tion energy released in light. ILEDs generally have great light
eciency, long lifetime, and environmental stability compared
to other light-emitting devices. However, intrinsic rigidity
from inorganic materials is a critical obstacle to morphological
change, therefore, proper integration methods with polymeric
substrates are highly required to overcome a significant mod-
ulus mismatch.
The island-bridge configuration is an early method to inte-
grate ILEDs with stretchable substrates.[139–141] The epitaxially
grown ILEDs on wafer inevitably require additional processes
to transfer them to other substrates. Park et al. utilized elas-
tomeric stamp with kinetic control of adhesion in order to
transfer ILEDs array mesh to a pre-strained PDMS substrate.
The Ti/SiO2 deposited island structures successfully incorpo-
rated with the elastomeric substrate. When the external strain
was applied to the PDMS, the chemically bonded islands were
not deformed, while the serpentine electrodes deformed out-
of-plane to accommodate the strain, ensuring a reliable opera-
tion of the stretchable ILEDs array under a uniaxial strain up to
22%. (Figure 13a).[139] Similarly, Byun etal. also systematically
fabricated strain-engineered stretchable platforms by adopting
inkjet-printed rigid islands (PRIs). PMMA ink was formulated
by changing its molecular weight and the number of printing
layers was optimized for the robustness of the islands. After
printing the PRIs, another top PDMS was covered to increase
stability, preventing the PRIs from being fractured under a uni-
axial strain of 40%. Then, various inorganic chips, including
ILEDs, integrated onto strain-free, engineered PRI employing
epoxy adhesives, enabling on-skin electronics. Moreover, the
inkjet printing modifies island patterns and interconnects
freely, enhancing the degree of freedom of stretchable circuit
formation. (Figure13b).[104]
The island-bridge structures eectively suppressed the
deformation around ILEDs. However, the abrupt dierence of
elastic modulus can induce significant strain concentration at
the interfaces, which causes delamination of devices and plastic
deformation of substrates. To overcome failure issues, Biswas
etal. embedded whole circuit into Ecoflex with very low Young’s
modulus without additional island structure. The whole cir-
cuit components including ILEDs array, stretchable electrodes,
and VIAs were fabricated on the PI substrates, and then a low
melting point solders were coated on the contact pads, where
ILEDs array would be transferred by pick-and-place assembly
techniques. The ILEDs circuit on the PI substrate was fully
embedded by Ecoflex over molding, and then the PI substrate
was partially perforated by electron cyclotron resonance plasma
etching, enabling the circuits to be stretchable. The embedded
circuits could be stretched up to 260% and showed high
deformability under various 3D guided shapes (Figure13c).[100]
The conventional bonding that utilizes metal soldering
or rigid adhesive with high elastic modulus ensures robust
bonding, but they are simultaneously vulnerable to mechan-
ical deformation. Therefore, Hwang et al. fabricated intrinsi-
cally stretchable anisotropic conductive films (S-ACF) that can
directly interconnect ILEDs and substrates. The polystyrene-
block-poly(ethylene-ran-butylene)-block-polystyrene-graft-
maleic anhydride (SEBS-g-MA) was used as a free-standing
and stretchable film. The periodically arranged metal particles
were embedded in the S-ACF for electrical connection. The
boding of ILEDs was acquired by modifying the bottom sur-
face wettability of ILEDs with 3-aminopropyltriethoxysilane
(APTES). The MA-NH2 amide bond with SEBS-g-MA template
enabled stretchable ILEDs array with high stretchability of 70%
(Figure13d).[142]
Several integration methods have been proposed to over-
come large modulus dierence in the heterogeneous interface
between rigid ILEDs and soft substrates. Although the integra-
tion concept was proved for a few ILEDs, the successiveness
and repeatability to integrate a large number of ILEDs and
other electronic devices were not identified. Moreover, since
the size of the ILEDs has minimized, a systematic integration,
including mechanical and electrical connection, applicable to
fine-pitch ILEDs should be further investigated.
5.2. Organic Light-Emitting Devices
Organic light-emitting devices composed of organic materials
with electroluminescent property are one of reliable candidates
for light-emitting components in stretchable electronics due
to its intrinsic softness with mechanical robustness[143–148] and
processability on deformable substrates with low-temperature
process.[149,150] However, the stretchable organic light-emitting
devices should simultaneously sustain multiple axial strains. In
this regard, various approaches have been reported to achieving
stretchable organic light-emitting devices: using wrinkled struc-
ture and serpentine electrodes, island-bridge structures, and
intrinsically stretchable materials.
Wrinkled structure has been introduced as a representative
method for the fabrication of stretchable devices. To implement
the wrinkled structure on elastomeric platform, the stretch-
able OLEDs were fabricated on pre-stretched elastomer such
as PDMS or PU. When pre-stretched platform was released,
directional wrinkles were implemented depending on the direc-
tion of applied prestrain. When external strain was applied
along the wrinkle direction, the structure on the display was
unfolded without physical damage. Using this structure, Yokota
et al. attached ultrathin OLEDs on pre-stretched elastomer
(Figure 14a).[151] Ultraflexible OLEDs with thin encapsulation
layers were fabricated on 1-µm-thick parylene films, showing
stable operation under ambient condition. After lamination
process on pre-stretched elastomeric substrate, the device
could maintain its electroluminescent performance even
after 1000 cycles with 60% strain. The encapsulated stretch-
able display with ultrathin structure with total thickness of
3 µm could be easily laminated on human skin without
degradation, thereby applied to wearable device which can visu-
alize data extracted from human body. Yin etal. expanded the
Adv. Mater. Technol. 2023, 2201067
2365709x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/admt.202201067 by Seoul National University, Wiley Online Library on [08/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com
© 2023 Wiley-VCH GmbH
2201067 (20 of 37)
www.advmattechnol.de
stretchability of devices to 2D direction with random-located
buckling networks on the surface of OLEDs (Figure 14b).[152]
After attaching ultraflexible OLEDs on 100% pre-stretched elas-
tomer with both 2D directions, the random networks of micro-
scale buckling structures were formed after releasing process.
Through this approach, OLEDs was operated without degrada-
tion even with 50% of 2D strain. On the other side, Jeong etal.
focused on the image distortion of stretchable OLEDs with
wrinkle structures (Figure14c).[153] Unlike the other approaches
for the stretchable OLEDs with wrinkled structures, the OLEDs
were directly fabricated on pre-stretched elastomers without
lamination of plastic substrate. After fabrication of OLEDs and
releasing process, the mean values of size of wrinkles were all
less than 8 µm and those micro-wrinkles demonstrated both
distortion-free pixels and high stretchability.
By implementing stretchable OLEDs on strain-engineered
platform with serpentine shaped stretchable electrodes, Lim
etal. reported another promising approach based on lamination
Figure 13. Stretchable ILEDs. a) Island-bridge-based stretchable ILEDs fabricated by transferring an ILED array mesh to a PDMS substrate. Reproduced
with permission.[139] Copyright 2009, American Association for the Advancement of Science. b) Customizable stretchable platforms, fabricated by inkjet
printing of PRIs and electric circuits, and epoxy bonding of ILEDs. Reproduced with permission.[104] Copyright 2017, Springer Nature. c) Stretchable
multilayered circuits fabricated by embedding lithographically fabricated electric circuits into Ecoflex. Reproduced with permission.[100] Copyright 2019,
Springer Nature. d) Stretchable ACFs as direct interfaces for the ILEDs and stretchable substrates. Reproduced with permission.[142] Copyright 2021,
American Association for the Advancement of Science.
Adv. Mater. Technol. 2023, 2201067
2365709x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/admt.202201067 by Seoul National University, Wiley Online Library on [08/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com
© 2023 Wiley-VCH GmbH
2201067 (21 of 37)
www.advmattechnol.de
of 3D island-bridge structures.[62] Micropillar arrays placed on
PDMS substrate were laminated with OLEDs on SU-8 sub-
strate, therefore pillars supported the OLED pixels (Figure14d).
In addition, OLED pixels were connected with the bridges with
serpentine electrodes. Under the external strain, the serpentine
electrodes were stretched while pillars attached with PDMS
pixels were maintained its spacing. Because the mechanical
strain was dissipated to the serpentine electrodes and PDMS
substrate, the OLED arrays were reliably stretched up to 35%
without any image distortion.
Although applying island-bridge structures on OLEDs is
an ecient method for implementation of reliable stretchable
electroluminescent devices, intrinsically stretchable OLEDs
with novel material designing have a great deal of attention.
Liu et al. suggested intrinsically stretchable AM organic light-
emitting electrochemical cell (OLEC) array (Figure 14e).[70]
AgNW-coated polyurethane acrylate (AgNW-PUA) was used
as anodes and cathodes, and optimized blend of light-emit-
ting materials was adopted for better stretchability. Fabricated
OLECs were stable at 30% strain and vertically connected with
stretchable polymer semiconductors to fabricate AMOLECs. To
further enhance stretchability, Kim etal. blended Triton X-100
surfactant with active layers (Figure 14f).[71] When the sur-
factant was blended with emissive materials, they interrupted
Figure 14. Stretchable organic light-emitting devices. a) Schematic of ultraflexible OLEDs with passivation layers. Reproduced with permission.[151]
Copyright 2016, American Association for the Advancement of Science. b) Images of 1D stretchable OLEDs at strains of 0-80%. Reproduced with per-
mission.[152] Copyright 2016, American Chemical Society. c) Schematic of imperceptible/macroscopic wrinkles in stretchable OLEDs, and demonstration
of distortion-free pixel arrays. Reproduced with permission.[153] Copyright 2020, Wiley-VCH. d) Cross-sectional and tilted-view SEM image of a stretch-
able substrate, and images of stretchable OLEDs with strains of 0–30%. Reproduced with permission.[62] Copyright 2020, American Chemical Society.
e) Schematic layout and optical image of stretchable OLEC pixels driven by stretchable transistors, and images of a single OLEC pixel with dierent
strains. Reproduced with permission.[70] Copyright 2020, Springer Nature. f) Images of relative luminance change in devices under various 1D strains
and 3D-deformed OLEDs. Reproduced with permission.[71] Copyright 2021, American Association for the Advancement of Science.
Adv. Mater. Technol. 2023, 2201067
2365709x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/admt.202201067 by Seoul National University, Wiley Online Library on [08/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com
© 2023 Wiley-VCH GmbH
2201067 (22 of 37)
www.advmattechnol.de
interchain reactions between emissive materials, thereby
making layers stretchable. In addition, when Triton X-100 was
mixed with PEDOT:PSS which was used as hole transport layer
(HTL), the surfactant hinders the interactions between PEDOT
and PSS, so the hole transport can be enhanced and Young’s
modulus is decreased. The fabricated intrinsically stretch-
able OLEDs maintained their performance at 40% strain, and
showed stable emission even at the strain of 80%.
With promising advantages of OLEDs such as mechan-
ical reliability and possibility of large area process, various
approaches were adopted for fabrication of stretchable OLEDs.
However, stretchable encapsulation which provides long-term
stable operations even under the mechanical strain is a chal-
lenge for commercialization. Furthermore, fabrication of large
area OLEDs arrays and linkage with stretchable driving circuit
should be next steps for stretchable OLED displays.
5.3. Metal-Halide Perovskites LEDs
Metal-halide perovskites (MHP) became promising light-emit-
ting materials since the first development of MHP LEDs in
2014.[154] The narrow emission linewidth (<20 nm) of MHP can
achieve ultrahigh color purity with the color gamut ratio >140%.
MHP is commonly composed of ABX3 forming a cubic struc-
ture, where A is an organic ammonium or an alkali metal ion,
B is a transition metal cation and X is a halide anion.
Controlling dimensionality of MHP from 3D, 2D, 1D to
0D with dimension decreased from micrometer-scale down
to nanometer-scale is the most eective approach to confine
the charge carriers and to increase the possibility of radiative
recombination. Hence, using 0D nanocrystals could achieve
highly ecient LEDs on a glass substrate with EQE boosted
>23.4%.[155] However, the rigid nature of MHP nanocrystals
makes it dicult for stretchable applications.
Introducing the buckling structure to the device with thin
substrates will significantly reduce the strain experienced by
the device, as the magnitude of the strain was proportional to
the thickness of the substrate.[156] The AgNW percolation net-
works were embedded in the PI substrate (1–2 µm) to form an
ultrathin transparent conductive electrode with excellent chem-
ical stability against organic solvents.[157] After multiple succes-
sive solution processes and Al electrode deposition, the sub-
strate was conformably attached to pre-strained 3M-VHB tape
to create a buckling structure with a small bending radius of
70 µm (Figure 15a). The stretchable MHP-based LEDs showed
a low turn-on voltage (Von) of 3.2 V (Figure15b), the maximum
luminance of 3187 cd m2 at 9 V, and can retain the device per-
formance with tensile strain up to 50% (Figure15c).
As an alternative to rigid MHP nanocrystals, methylammo-
nium lead bromide (MAPbBr3) polycrystalline precursors are
mixed with an ion-conducting polymer with viscoelastic prop-
erties such as poly(ethylene oxide) (PEO) to reduce Young’s
modulus.[158] By mixing the conducting polymer PEDOT:PSS
with 33 wt% of PEO, the highest conductivity of 35600 S m1
was achieved and showed no significant change in conduc-
tivity under 20% of strain. The cathode was deposited on the
composite emitter using eGaIn (Figure 15d). The stretchable
MHP LEDs showed Von= 2.4 V (Figure15e), and the maximum
luminance of 15960 cd m2 at 8.5 V (Figure 15f ). A similar
approach has been made using CsPbBr3 with the addition of
PEO and poly(vinylpyrrolidone) (PVP).[159] The morphological
analysis on the failure mechanism of the stretchable MHP
LEDs showed that crack initiated at the perovskite layer and
prorogated to the bottom electrode.[160] Therefore, improving
the stretchability of the perovskite layer is the key for stretch-
able MHP LEDs.
Alternatively, an air-stable stretchable display has been
achieved by the integration of the perovskite stretchable color
conversion layer (SCCL) and the stretchable light-emitting
devices.[161] Perovskite nanocrystals were dispersed in the elas-
tomer due to the high compatibility between the ligand and
the polymer matrix. The SCCL down-converts the blue emis-
sion from stretchable light-emitting devices and reemits it as
green (Figure 15g). The PL intensity decreased by 29.8% at
100% strain and recovered to the initial state after releasing
(Figure15h). The stretchable light-emitting devices with SCCL
show the stretchability of 180% (Figure15i).
Despite approaches have been made to apply organic-
inorganic hybrid perovskites for stretchable applications, more
eorts are needed to increase both stretchability and device
light-emitting eciency with innovations in materials aspect.
5.4. ACEL Devices
ACEL devices are normally operated based on hot-electron
impact on the light-emitting phosphors under a high electric
field across the thick dielectric layer >10 µm. Phosphors are
embedded in the stretchable dielectric layer with a high dielec-
tric constant to reduce potential drop, and the dielectric layer is
sandwiched between two stretchable electrodes. As there is no
direct charge injection from the electrode to the phosphors, the
conductivity of the electrode for the ACEL devices is lower than
that of other LEDs.
To increase the stretchability of the ACEL devices, an ionic
conductive hydrogel is incorporated as a stretchable electrode.
With the application of voltage, charges in the electrode are
separated at the contact/hydrogel interface over nanometers,
which can create large capacitance on the order of 101 F m2.[162]
Hygroscopic lithium chloride was adopted as the ionic con-
ductor with high conductivity of 10 S m1, while stretchable
polyacrylamide was used as the elastomer matrix.[163] The low
elastic modulus of the hydrogel allows the free stretching of
the device without delamination at the interface (maximum
strain 480%). The device can be operated at 700 Hz under
25 kV cm1 with a luminous ecacy of 43.2 mlm W1. The
matrix with an 8 × 8 array of 4-mm pixels has been fabricated
using the replica molding technique, which can also undergo
stretching, rolling, folding, and wrapping.
As an alternative to the molding technique, a stretchable
multicolor display can be prepared using the transfer printing
method. Phosphors mixed with inks can be photopatterned
first and then dry transferred to the dielectric layer using a
thermal release tape. Multiple cycles of transfer printing allow
the assembly of multicolor pixels on the large dielectric.[164]
Adv. Mater. Technol. 2023, 2201067
2365709x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/admt.202201067 by Seoul National University, Wiley Online Library on [08/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com
© 2023 Wiley-VCH GmbH
2201067 (23 of 37)
www.advmattechnol.de
The ionic conductive hydrogel can also be constructed in
the fiber shape to further boost the stretchability up to 800%
(Figure 16a).[165] Two ellipse-shaped fibers wrapped by ZnS
phosphors and elastomer composite to display programmable
pattern. The electroluminescent layer also serves as the pro-
tecting layer for the hydrogel electrode. The stretchable light-
emitting fiber was woven into a stretchable textile display. Also,
owing to the high stretchability of hydrogel that can sustain
repetitive stretching to 700% for 1000 cycles, another ACEL
device can undergo tensile strain up to 700% with maximum
luminance = 95 cd m2 (Figure16b).[166]
Dierent from the monotonic increase in light emission of
ionic conductive hydrogel-based ACEL under stretching, the
percolation AgNW-based electrode showed a slight increase
in emission intensity of 30% followed by a monotonic
decrease in intensity with further stretching (Figure 16c).[167]
The electric field concentrated on the AgNW decreased with
a larger open area created between AgNWs under stretching,
resulting in a decrease in the light intensity.[166] Hence, main-
taining the high electric field under large deformation is greatly
challenging especially for conventional transparent conductive
electrodes based on low-dimensional materials and conducting
polymers. Incorporation of dielectric particles such as BaTiO3
can eectively enhance the electric field inside the stretch-
able light-emitting layer, which substantially increases the
electroluminescent intensity (Figure16d).[168]
In addition, the high operating voltage of stretchable ACEL
devices is another issue that limits practical applications; this
is caused by the low permittivity of the elastomer. Hence,
approaches have been made to increase the permittivity of the
Figure 15. Stretchable perovskite LEDs. a) Device structure, b) current density (I)–voltage (V)–luminance (L), and c) digital images of geometrically
stretchable perovskite quantum dot LEDs. Reproduced with permission.[157] Copyright 2019, Wiley-VCH. d) Digital images, and e) IV and f) LV char-
acteristics, of intrinsically stretchable perovskite LEDs. Reproduced with permission.[158] Copyright 2017, Wiley-VCH. g) Device structure and digital
images of stretchable LEDs before and after integration with perovskite stretchable color conversion layer (SCCL). h) PL spectra of the SCCL under
uniaxial tensile strain. i) Digital images of the stretchable LED integrated with the SCCL under tensile strains of up to 180%. Reproduced with permis-
sion.[161] Copyright 2020, Wiley-VCH.
Adv. Mater. Technol. 2023, 2201067
2365709x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/admt.202201067 by Seoul National University, Wiley Online Library on [08/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com
© 2023 Wiley-VCH GmbH
2201067 (24 of 37)
www.advmattechnol.de
dielectrics. Poly(vinylidene fluoride)-based fluoroelastomer has
also been developed as the dielectric materials that can main-
tain high permittivity and self-healing properties with the addi-
tion of non-ionic additives (Figure16e). The ACEL has achieved
a high brightness of 1460 cd m2 at 2.5 V µm1 with a maximum
stretchability of 800%.[169] Besides adding BaTiO3 particles into
the light-emitting layer will also lead to higher permittivity and
brightness (Figure16f).[170] With above merits, stretchable ACEL
devices have been successfully demonstrated for the electronic
skin that is composed of strain and electrocardiogram sensors.
Given the fact that the goal of the stretchable ACEL display is
the wearable application, both maximum luminance and device
eciency need to be improved to meet the requirement of low
power consumption. More innovations in stretchable light-
emitting materials are needed to further boost the eciency.
6. Stretchable LED Matrix Arrays
In the display, multiple LED units that compose a matrix array
are driven individually. Stretchable LED arrays in which geo-
metrically or intrinsically stretchable structures, materials, and
electronic devices (transistors and LEDs) technologies are inte-
grated are manufactured and driven in PM or AM structures.
In this chapter, the structure and characteristics of stretch-
able displays according to 1) PM and 2) AM operations, and 3)
high-resolution full-color displays at the prototype level will be
discussed.
6.1. PM Arrays
PM LED array, consisting of driver integrated circuit chip and
light-emitting units, is operated by turning on the light-emitting
units line by line, where each pixel is defined at the overlapped
points of driving electrodes. Thus, PM structure can be obtained
by simple manufacturing process without complicated circuit
design. However, PM structure has limitations in increasing
resolution or panel size due to large power consumption.[171]
So, it has been utilized in limited areas, such as low-end and
small-sized mobile applications. In this regard, stretchable PM
array has been demonstrated to show 1) a novel unit technology
for integrated stretchable systems, including stretchable light-
emitting units, interconnects, and driving circuit chips,[30,172–175]
Figure 16. Stretchable ACEL devices. a) Digital images of a stretchable light-emitting fiber being stretched from 0% to 800%. Reproduced with permis-
sion.[165] Copyright 2018, Wiley-VCH. b) Digital images of stretchable ACEL devices stretched with dierent strains. Reproduced with permission.[166]
Copyright 2016, Wiley-VCH. c) Device structure of ACEL device using AgNW as the electrode. Reproduced with permission.[167] Copyright 2015, Wiley-
VCH. d) Schematic of ACEL device that incorporates BaTiO3 particles. Reproduced with permission.[168] Copyright 2016, Wiley-VCH. e) Molecular
structure of a fluoroelastomer with a non-ionic fluorinated surfactant for a healable and low-field illuminating optoelectronic stretchable device. Repro-
duced with permission.[169] Copyright 2019, Springer Nature. f) Multifunctional self-healable electronic skin device on skin with ACEL devices emitting
a blue-green light. Reproduced with permission.[170] Copyright 2019, Springer Nature.
Adv. Mater. Technol. 2023, 2201067
2365709x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/admt.202201067 by Seoul National University, Wiley Online Library on [08/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com
© 2023 Wiley-VCH GmbH
2201067 (25 of 37)
www.advmattechnol.de
or 2) the concept of next-generation wearable displays, such as
skin-attachable patch-type display and textile display integrated
with health monitoring sensors.[11,104,176–178]
To demonstrate stretchable PM LED array, island-bridge
structure with stretchable interconnects is a practical approach.
In island-bridge structure, light-emitting units are located at
rigid-island region, where the strain is suppressed by structural
design, and they are interconnected by stretchable electrodes in
bridge region, where the strain is released. Therefore, inherent
stretchability of light-emitting unit become less essential in the
island-bridge structure, enabling ILED to be utilized in stretch-
able PM LED array.[30,172–176] Thus, several studies have been
implemented ILED-based PM array to demonstrate the feasi-
bility of novel structural designs such as serpentines,[172] helical
structures,[174] and intrinsically stretchable materials.[173]
Meanwhile, some research focused on fill factor of stretch-
able display beyond the stretchability itself.[30,175] Increasing
stretchability has trade-o relationship between the fill factor
of LEDs, deteriorating the resolution of display. Therefore, a
double-layered modular design separating LED unit substrates
and stretchable interconnectors (Figure 17a)[30] or 3D assembly
with hidden pixel structure[175] has been suggested. Meanwhile,
Byun etal. demonstrated a PM LED array on inkjet printing-
based stretchable platform, including printed rigid island
and wrinkled Ag interconnects in a fully-customizable way
(Figure 17b).[104] By integrating ILED chips, microcontroller
units, and other surface-mount device components on a rigid
island-embedded stretchable platform, a 2-bit digital multiplier
and temperature monitoring system were demonstrated on
skin.
Stretchable PM LED array with island-bridge structures can
also be made with thin-film-based LEDs such as OLED and
quantum dot LED (QLED) to make them more flexible and con-
formal.[11,177] Therefore, real-time health monitoring patch that
provided biological information on human skin in intuitive way
has been reported. Lee etal. demonstrated integrated system of
health monitoring patch with stretchable OLED array, photop-
lethysmography (PPG) heart rate sensor, processing modules,
and batteries (Figure17c).[11] Stretchable OLED array could be
realized by modulus-engineered strain relief layer and micro-
cracked interconnects, making compatible with lithography
processes. Lee et al. also demonstrated a sensor-integrated
Figure 17. Stretchable PM LED array. a) Stretchable PM ILED array with high fill factor. Reproduced with permission.[30] Copyright 2022, American
Chemical Society. b) Customizable inkjet printing-based stretchable PM ILED array. Reproduced with permission.[104] Copyright 2017, Springer Nature.
c) Real-time health monitoring patch with stretchable OLED array and PPG sensor. Reproduced with permission.[11] Copyright 2021, American Associa-
tion for the Advancement of Science. d) Stretchable QLED array for monitoring body movements and temperature. Reproduced with permission.[177]
Copyright 2022, Elsevier. e) Textile-based functional systems with intrinsically stretchable ACEL fibers. Reproduced with permission.[178] Copyright 2021,
Springer Nature.
Adv. Mater. Technol. 2023, 2201067
2365709x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/admt.202201067 by Seoul National University, Wiley Online Library on [08/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com
© 2023 Wiley-VCH GmbH
2201067 (26 of 37)
www.advmattechnol.de
stretchable QLED array to monitor body movement and skin
temperature (Figure 17d).[177] QLED array deposited on NOA
63 islands were operated by stretchable LM interconnects.
Beyond island-bridge structure, fully stretchable PM array
were demonstrated by using intrinsically stretchable ACEL
device (Figure17e).[178] By weaving stretchable conductive fiber
and ZnS-coated conductive fiber, each electroluminescent
pixel could be defined at contact points. While they require AC
operation with high voltage, the feasibility of textile electronics
integrated with displays and other electronic components has
proven to broaden the applicable areas of stretchable elec-
tronics and more intuitive human-machine interfaces. Various
next-generation stretchable technologies are expected to con-
tinuously be demonstrated through PM driving scheme, sug-
gesting novel concept of wearable display.
6.2. AM Arrays
PM has a simpler structure and more straightforward operation
than AM by applying voltage to each data and scan line sequen-
tially to emit light at intersecting pixels. PM divides the scan
time by the number of lines and sequentially applies voltage
only for a short period of time. During one frame scan, only
pixels to which voltages of the scan line and the data line are
applied simultaneously emit light. Therefore, a high voltage is
required to ensure high brightness during a short light emis-
sion time, which shortens the lifespan of the materials. In addi-
tion, as the number of lines increases, more time division is
required, so there is a limitation in increasing resolution and
size.
AM has transistors and storage capacitors in each pixel, so
it is possible to store voltage for a certain period of time and
maintain light emission for one frame. Also pixels are individu-
ally controlled using transistors of the backplane. Therefore, it
is suitable for high-resolution large-area displays with advan-
tages such as low-power operation, minimization of line cross-
talk and flicker, fast response, and capable of compensation
circuit configuration. The AM stretchable display is more chal-
lenging than the PM stretchable display because the mechan-
ical stretchability of backplane circuits and their components
(e.g., transistors and capacitors) should be improved. The devel-
opment of AM stretchable display is still in the early stages and
many breakthrough studies are required, but it is essential to
produce a future high-resolution large-area stretchable display.
A few stretchable AM LED arrays have been reported with the
extrinsically stretchable geometries and rigid inorganic/organic
LEDs or the intrinsically stretchable geometries and polymer
LECs. Also, the mature high-mobility silicon transistors or
the next-generation transistors that exploit oxide-, CNT-, or
polymer–semiconductors have been used for driving AM LED
arrays.
A 32 × 32 stretchable AM ILED array (13 ppi) was fabricated
on an island-bridge patterned PI substrate with Cu serpentine
interconnects.[179] A 2T-1C structure backplane composed of
amorphous IGZO TFTs with a mobility 10 cm2 V1 s1 drives
rigid ILED pixels. An LED pixel array was connected to a pix-
elated backplane with isotropic conductive adhesive and pick-
and-place technique, then encapsulated with a thin TPU film.
The LED chips array had uniform brightness, and showed flex-
ibility and conformability on the complex surfaces without sig-
nificant loss in the uniformity.
μ-LED refers to an ultrasmall LED with a size of less than
100 µm, and is attracting attention as a next-generation display
based on high brightness, low power, high resolution, infi-
nite contrast, fast response, wide color gamut, long lifespan,
and stability in air and moisture. Also, fast-response and low-
power single crystal Si-TFT with the high field-eect mobility
(700 cm2 V1 s1) which is a desirable candidate for large-area
and high-resolution display was used to drive μ-LED array.
Stretchable AM array composed of 8 × 8 μ-LED and Si-TFT
array was fabricated on the PDMS substrate using roll transfer
printing (Figure 18a).[141] High mobility Si-TFTs and μ-LEDs
were fabricated separately and sequentially transferred to a tem-
porary glass substrate with a high yield and precise alignment.
Then the Si-TFTs and μ-LEDs are connected by serpentine
interconnects and transferred on a rubber substrate to demon-
strate stretchable AM μ-LED array. The operation of μ-LED by
Si TFT was controlled with high on-o current ratio (107). The
rigid island part was firmly fixed on the PDMS substrate, and
the serpentine interconnect at the neutral mechanical plane
composed of metal layer sandwiched by polymer was slid on
the PDMS surface to maintain the light emitting properties
without deterioration even at 40% tensile strain (Figure18b,c).
Especially, the strain applied to the rigid island area was less
than 0.1%, so the active components maintained the stable
characteristics even after 200 stretching cycles at 40% strain.[141]
An island-bridge structure using intrinsically stretchable
and straight interconnects instead of the serpentine intercon-
nects can minimize the space occupied by the interconnects
and thereby potentially increase the pixel density and aperture
ratio. LM that exhibits small resistance change under strain
can be a possible candidate instead of serpentine patterned
flexible metal interconnects. SWCNT transistors and μ-LED
arrays were fabricated on rigid polyethylene terephthalate (PET)
islands embedded in Ecoflex substrates, and each island was
connected by intrinsically stretchable and straight LM intercon-
nects (Figure18d).[180] In bending or stretching, the strain was
concentrated in the Ecoflex region with low modulus (69 kPa),
and the negligible strain was applied in the rigid PET island
region with high modulus (2.0–2.7 GPa). For example, at a
biaxial strain of 30%, Ecoflex is subjected to 240% strain, while
PET is subjected to near 0% strain. Therefore, the 5 × 5 array
SWCNT TFTs fabricated in the PET region shows a drain
current change less than 4% at 30% biaxial strain. As a result,
5 × 5 AM μ-LED array fabricated with 10 μ-LEDs on each PET
island did not show any deterioration in brightness under 30%
biaxial strain and even after 1000 bending cycles (Figure18e).
Compared to the geometrical strain engineering
approaches, intrinsically stretchable electronic components
can increase stretchability of AM LED array by reducing the
portion of nonstretchable area and might be better for the
electronic skin devices with smaller mechanical constraints on
the skin. AMOLEC array integrated with intrinsically stretch-
able organic TFTs and intrinsically stretchable OLECs main-
tained brightness without deterioration at 30% tensile strain
(Figure 18f).[70] In stretchable organic TFTs, the high mod-
ulus (>2.25 GPa) of the isoindigo moiety-containing polymer
Adv. Mater. Technol. 2023, 2201067
2365709x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/admt.202201067 by Seoul National University, Wiley Online Library on [08/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com
© 2023 Wiley-VCH GmbH
2201067 (27 of 37)
www.advmattechnol.de
semiconductor, which cracks at 15% tensile strain was low-
ered to <0.75 GPa by blending an azide functionalized PDMS
crosslinker. The crosslinked semiconductor film did not show
cracks even at 100% tensile strain. In addition, crosslinkable
stretchable insulators that have the modified hydroxyl groups
of perfluoropolyether diols with dimethacrylate groups showed
very negligible change in dielectric constant at 100% strain.
With CNT stretchable gate, source and drain electrodes, the
5 × 5 transistor array showed less than one order of magnitude
decrease in both the mobility and on-current, and minimal
changes in leakage current even after 1000 repeated stretching
at 100% strain.
In stretchable OLECs, the AgNWs-PUA electrode and
PEDOT:PSS hole injection layer sandwiched the light emission
layer that is composed of a blend of a light-emitting polymer
(Super Yellow), an ion-conducting polymer (ethoxylated tri-
methylopropane triacrylate), and lithium trifluoromethane sul-
fonate, which enables formation of PIN junction in the light
emission layer. The stretchable OLECs did not showed delami-
nation, crack formation or degradation of current density under
30% strain. An intrinsically stretchable 2 × 3 AMOLEC array
showed stable operation in bending, twisting, and stretching
states (Figure18g).[70]
Based on such stretchable PM and AM LED arrays intro-
duced in this chapter, a large-area, high-resolution, and full-
color stretchable display at a level closer to commercialization
will be introduced in the next chapter.
6.3. High-Resolution Full-Color Displays
The development of high-resolution, high-density and full-
color stretchable displays is mainly conducted by display
industry because of the technological manufacturing maturity
and capacity of fabrication facilities. Although many prom-
ising intrinsically stretchable light emitting devices intro-
duced in the previous chapters are being actively developed,
important factors for manufacturers and users such as pro-
cessing yield, reliability, uniformity, eciency, lifespan, and
outdoor visibility must be considered to develop commercial
products. Therefore, in the current stage, the island-bridge
structure with established non-stretchable LEDs and transis-
tors is a viable approach for developing a prototype stretchable
displays.
PM stretchable LED array with 10 × 10 RGB LED package
chips and serpentine copper interconnects on the PI substrate
was developed in 2015 and demonstrated stable operation
at 10% strain.[172] It proved the scalability by demonstrating
80 × 45 array PM LED display with 8.5 ppi. Also recently, a 2.7 in.
and 42 ppi (96 × 60) stretchable display with 90 × 150 µm RGB
LED chips was reported.[12] The LED chips are transferred by
pick-and-place, stamp, and self-assembly techniques on the
island-bridge-patterned flexible backplane (Figure 19a). Then,
assembled backplane was subsequently transferred to and
encapsulated by low modulus elastomer that are crosslinkable
at low temperature and highly transparent.
Figure 18. Stretchable AM LED array. a) Schematic and digital image of stretchable AM μ-LED array with single crystal Si-TFT backplane. b) Digital
and c) optical images of stretchable AM μ-LED array with serpentine interconnects under strains of 0% and 40%. Reproduced with permission.[141]
Copyright 2017, Wiley-VCH. d) Digital image and schematic of AM μ-LED array with SWCNT TFTs. e) Digital images of stretchable AM μ-LED array
with intrinsically stretchable straight LM interconnects under strains of 0% and 30%. Reproduced with permission.[180] Copyright 2016, Wiley-VCH.
f) Schematic and g) digital images of stretchable AMOLEC array that incorporates intrinsically stretchable OTFT and OLEC arrays. Reproduced with
permission.[70] Copyright 2019, Springer Nature.
Adv. Mater. Technol. 2023, 2201067
2365709x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/admt.202201067 by Seoul National University, Wiley Online Library on [08/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com
© 2023 Wiley-VCH GmbH
2201067 (28 of 37)
www.advmattechnol.de
The stretchability of the panel was simulated in terms of
the ratio between island and bridge parts, the mechanical
modulus of elastomer, and the width of interconnect and flex-
ible substrate. To achieve high stretchability of panel, smaller
island/bridge ratio, smaller width of interconnect, and lower
modulus elastomer are preferred. The thickness of intercon-
nects and flexible substrate have negligible eects. With 600
µm pixel pitch, the fabricated 42 ppi stretchable display panel
demonstrated stable operation with biaxial stretching, free-from
twisting, and convex deformation by poking (Figure 19b-d).
This approach has a potential to be extended to 120 ppi stretch-
able display by employing 200 µm pixel pitch and μ-LED with
smaller feature size of 20 µm.
AM A 9.1 in. full-color stretchable AMOLED display was
demonstrated with island-bridge structure (Figure 19e).[9] The
AM display exploits an OLED array which can be directly fab-
ricate on the flexible substrate, and a flexible low-temperature
polycrystalline silicon (LTPS) backplane that exhibits excellent
properties for the high resolution and high frame rate display
with higher carrier mobility (50–100 cm2 V1 s1) than a-Si
(1 cm2 V1 s1), organic (1 cm2 V1 s1), and metal oxide semi-
conductors (10 cm2 V1 s1). The electrical stability of LTPS TFTs
is usually suered from mechanical deformations because
both interface state density and grain boundary state density
increase under bending and compression states. The induced
instability of leakage current and threshold voltage changes
charges stored in the capacitor, resulting in nonuniform bright-
ness of pixel array. The strain applied to the LTPS TFTs was
minimized by the geometrical approach of rigid island and ser-
pentine bridge structure under 5% tensile strain. Therefore, the
Figure 19. Stretchable high-density full color display. a) Schematics of fabrication process for a stretchable PM ILED display. Digital images of a stretch-
able PM ILED display under various deformations, including b) biaxial stretching, c) twisting, and d) poking. Reproduced with permission.[12] Copyright
2021, Wiley-VCH. Digital images of a 9.1 in. stretchable AM OLED display e) without deformation and with f) convex and g) concave deformations.
Reproduced with permission.[9] Copyright 2017, Wiley-VCH. h) Digital image of a 14.1 in. stretchable AM OLED with selective z-axis stretching marked
with a white arrow at #1 and without deformation at #2 and #3. Reproduced with permission.[29] Copyright 2019, Wiley-VCH.
Adv. Mater. Technol. 2023, 2201067
2365709x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/admt.202201067 by Seoul National University, Wiley Online Library on [08/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com
© 2023 Wiley-VCH GmbH
2201067 (29 of 37)
www.advmattechnol.de
TFTs maintained stable electrical characteristics of which the
threshold voltage shift was lower than 0.1 V and the leakage
current was reduced due to aging eect.
The mechanical modulus of the flexible backplane on the
patterned PI substrate was 40 MPa which is in the similar range
of soft rubber and is much lower than the modulus of PI film
(4.2 GPa). Convex and concave shaped displays with 10 mm ver-
tical displacement are demonstrated through thermoforming
process by attaching the deformed displays on the patterned
mold with thermoplastic sheet and adhesive (Figure 19f,g).
Moreover, a 14.1 in. full-color stretchable AMOLED display
was developed and it demonstrated no significant degradation
under deformation with 45 mm of vertical height and even after
10000 stretching cycles with 5% strain (Figure19h).[29]
7. Conclusion and Outlook
Stretchable displays, which can be freely bent in any direction
and fit on any shape, are almost the end of the form factor
innovation beyond flexible displays. Stretchable displays will
maximize the portability of users’ devices with multidirectional
folding, enhance bodily comfort with skin-level moduli and
softness, and enhance versatility through multi-curvature defor-
mation. In the case of flexible displays, repeated deformation
only occurs in limited bent areas and typically involves unidi-
rectional folding. Moreover, flexible displays with bendable,
foldable, and rollable form factors can minimize the actual
elongation of active devices by placing them in a mechanically
neutral plane. In contrast, stretchable displays involve elonga-
tion in length, which leads to the development of stretchable
geometries and soft materials with stable electrical and optical
properties under elongation to give displays a free-form factor.
This paper reviewed the current state of the research on
structural designs and elastic materials with important features,
including the conductors for interconnections, semiconductors
for TFTs, and light-emitting materials for stretchable displays
(Figure 20). Furthermore, stretchable display prototypes that
have already been demonstrated based on the aforementioned
technologies were discussed.
The various approaches that are actively being studied to
develop a stretchable display all have their pros and cons. For
example, as mentioned above, geometrically stretchable struc-
tures such as buckled structures and island-bridge structures,
which are close to flexible, have the advantage that the existing
semiconductor/display devices and manufacturing processes
can be used. However, buckled structures are limited to pro-
ducing highly integrated arrays with high stretchability, and
island-bridge structures, which only increase the gap between
pixels, can produce severe image distortion. Intrinsically
stretchable structures can control the deformation of the pixel
area by adjusting the Young’s modulus. Thus, they can pro-
duce more natural image deformation and minimize the het-
erogeneity on the skin, but the development of materials with
high performance, long-term stability, and high reliability is
required.
For the commercialization of stretchable displays, the required
specifications can be varied according to their target applications
such as IoT devices, wearables, and automobiles. Nonetheless,
three key design parameters have been identified. First, the
stretchability of the display should be higher than 30%, which
approximately corresponds to a bending radius of 0.06 mm.
Stretchable displays attached to human skin should be mechani-
cally robust and operate stably under at least 30% strain, taking
into consideration the strain variation on the skin.[6,11,181,182]
Second, the resolution of the display should be higher than
200 ppi, accounting for the viewing distance and size of the
display. Third, the display should have a high brightness at low
operation voltages for outdoor visibility (e.g., 1000 cd m2 at 5 V).
Consequently, for high-resolution stretchable displays, the
conductivity of the stretchable conductor should be higher
than 5 × 106 S m1 to prevent a voltage drop across the narrow
interconnect lines, which causes nonuniformity of the dis-
play. Considering the conductivity and the change in resist-
ance under strain, metal-based stretchable conductors such as
LMs, metal thin films, and metal nanomaterials are promising
stretchable interconnects. LM has excellent deformability and
decent conductivity (3.5 × 106 S m1),[183–185] but it is neces-
sary to completely control the stability and flowability in the
subsequent lamination process. In addition, stability must be
ensured when the electronic devices are exposed to tempera-
tures below the melting points (e.g., 15.5°C for eGaIn, –19 °C
for Galinstan) of LMs.[183–185] For metal thin films, deformability
can be achieved by introducing a serpentine structure in addi-
tion to having excellent conductivity. However, serpentine inter-
connects are expected to be much longer and narrower than
straight interconnects, and metals with higher conductivities
than those of conventional interconnects are required.[186] Metal
nanomaterials such as AgNWs are promising next-generation
stretchable interconnects because of their excellent conductivity
and deformability. However, to apply them to mass production,
it is necessary to achieve high uniformity over a large area (e.g.,
6th generation mother glass, 1500 × 1850 mm).
Figure 20. Strategies to develop materials and devices toward realization
of stretchable display.
Adv. Mater. Technol. 2023, 2201067
2365709x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/admt.202201067 by Seoul National University, Wiley Online Library on [08/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com
© 2023 Wiley-VCH GmbH
2201067 (30 of 37)
www.advmattechnol.de
For the operation of OLED pixels, driving and switching
transistors require high charge carrier mobility >10 cm2 V1
s1, low o-current <10 pA, large on/o current ratio (>106),
and low hysteresis of threshold voltage <0.1 V.[187,188] In addi-
tion, it may be necessary to analyze the distribution of strain
required in the device in order to optimize the arrangement
of the stretchable and rigid parts, so that the stretchable dis-
play and the rigid driving circuit and power source can be
integrated. For example, the strain on skin near the wrist is
approximately 30%, which is much higher than the strain on
skin further from the wrist (<7%).[6,11] In addition, rigid parts
are required for the protection and handling of devices in daily
life, and rigid circuits and batteries can be integrated into hard
parts. Flexible and stretchable driving circuits and batteries are
also being actively researched.[85,102,194,106,135,136,189–193] This will
enable the realization of all-parts-flexible/stretchable electronic
devices. Additionally, the user should ideally be able to control
the deformation of the devices in the desired situation. For this
purpose, controlling the mechanical properties of the substrate
using materials such as shape-morphing materials can be a
suitable approach.[195–200]
Despite many research achievements, the specific form and
applications of stretchable displays are still relatively vague.
Recent studies developed deformable displays that could over-
come the design constraints and be applied to the A-pillar
and dashboard of an automobile, a health monitoring device
attached to the skin, and smart fiber-based clothing. Thus, inno-
vation activities to support various user scenarios and overcome
technical obstacles are becoming more common. Further work
still remains, including work to improve the display resolution,
stretchability, and reliability, but it is expected that they will be
on the market in the near future as next-generation form-factor
displays.
Acknowledgements
Y.L. and H.C. contributed equally to this work. This work was supported
by the Industry technology R&D program (20010427) funded By the
Ministry of Trade, Industry & Energy (MOTIE, Korea). This work was also
supported by the National Research Foundation of Korea (NRF) grant
funded by the Korea government (Ministry of Science and ICT) (NRF-
2016R1A3B1908431) and the Creative-Pioneering Researchers Program
through Seoul National University (SNU). This work was also supported
by Basic Science Research Program through the National Research
Foundation of Korea (NRF) funded by the Ministry of Education
(NRF-2021R1A6A3A03038934).
Conflict of Interest
The authors declare no conflict of interest.
Keywords
flexible displays, stretchable electronics, stretchable LEDs, stretchable
materials, stretchable transistors, wearable displays
Received: July 1, 2022
Revised: October 4, 2022
Published online:
[1] M. H.Miraz, M. Ali, P. S.Excell, R.Picking, in 2015 Internet Tech-
nologies and Applications, IEEE, Piscataway, NJ 2015, pp. 219–224.
[2] S.Nahavandi, Sustainability 2019, 11, 4371.
[3] A.Shcherbina, C.Mattsson, D. Waggott, H.Salisbury, J. Christle,
T.Hastie, M.Wheeler, E.Ashley, J. Pers. Med. 2017, 7, 3.
[4] B.Bent, B. A. Goldstein, W. A.Kibbe, J. P.Dunn, npj Digital Med.
2020, 3, 18.
[5] M.Etiwy, Z.Akhrass, L.Gillinov, A.Alashi, R.Wang, G.Blackburn,
S. M. Gillinov, D. Phelan, A. M. Gillinov, P. L. Houghtaling,
H.Javadikasgari, M. Y.Desai, Cardiovasc. Diagn. Ther. 2019, 9, 262.
[6] G. H.Lee, H.Kang, J. W.Chung, Y.Lee, H.Yoo, S.Jeong, H.Cho,
J.-Y. Kim, S.-G. Kang, J. Y. Jung, S. G. Hahm, J. Lee, I.-J. Jeong,
M.Park, G. Park, I. H. Yun, J. Y.Kim, Y.Hong, Y. Yun, S.-H.Kim,
B. K.Choi, Sci. Adv. 2022, 8, eabm3622.
[7] S.-C.Jo, J.-H.Hong, Dig. Tech. Pap. - Soc. Inf. Disp. Int. Symp. 2021,
52, 741.
[8] J.-Y.Yan, J.-C.Ho, J.Chen, Inf. Disp. 2015, 31, 12.
[9] J.-H.Hong, J. M.Shin, G. M.Kim, H.Joo, G. S.Park, I. B.Hwang,
M. W.Kim, W.-S.Park, H. Y.Chu, S.Kim, J. Soc. Inf. Disp. 2017, 25,
194.
[10] Q.Yang, P.Wang, F.Cao, L.Liu, Z.Liu, S.Shi, D.Wang, Dig. Tech.
Pap. - Soc. Inf. Disp. Int. Symp. 2020, 51, 1142.
[11] Y.Lee, J. W.Chung, G. H. Lee, H.Kang, J.-Y.Kim, C. Bae, H.Yoo,
S. Jeong, H. Cho, S.-G. Kang, J. Y. Jung, D.-W. Lee, S. Gam,
S. G. Hahm, Y. Kuzumoto, S. J. Kim, Z. Bao, Y. Hong, Y. Yun,
S.Kim, Sci. Adv. 2021, 7, eabg9180.
[12] J. Kang, H. Luo, W. Tang, J. Zhao, Y.-M. Wang, T.Tsong, P.Lu,
A. Gupta, L. Zeng, Z. Zhang, J. Zhou, S. Wang, R.Ma, X.Chen,
B.-G.Lee, Z.Yuan, P.Wei, X.Yu, Dig. Tech. Pap. - Soc. Inf. Disp. Int.
Symp. 2021, 52, 1056.
[13] N.Matsuhisa, X.Chen, Z. Bao, T.Someya, Chem. Soc. Rev. 2019,
48, 2946.
[14] L.Yin, J.Lv, J.Wang, Adv. Mater. Technol. 2020, 5, 2000694.
[15] X.Hu, Y.Dou, J.Li, Z.Liu, Small 2019, 15, 1804805.
[16] D.-Y. Khang, H. Jiang, Y.Huang, J. A. Rogers, Science 2006, 311,
208.
[17] D.-H. Kim, J.-H. Ahn, W. M. Choi, H.-S. Kim, T.-H.Kim, J. Song,
Y. Y. Huang, Z. Liu, C. Lu, J. A. Rogers, Science 2008, 320,
507.
[18] Y.Sun, W. M.Choi, H.Jiang, Y. Y.Huang, J. A.Rogers, Nat. Nano-
technol. 2006, 1, 201.
[19] M. Drack, I. Graz, T. Sekitani, T. Someya, M. Kaltenbrunner,
S.Bauer, Adv. Mater. 2015, 27, 34.
[20] M. Kaltenbrunner, T. Sekitani, J. Reeder, T. Yokota, K. Kuribara,
T.Tokuhara, M.Drack, R.Schwödiauer, I.Graz, S.Bauer-Gogonea,
S.Bauer, T.Someya, Nature 2013, 499, 458.
[21] M. Kaltenbrunner, M. S. White, E. D. Głowacki, T. Sekitani,
T. Someya, N. S. Sariciftci, S. Bauer, Nat. Commun. 2012,
3, 770.
[22] D. W.Kim, M.Kong, U.Jeong, Adv. Sci. 2021, 8, 2004170.
[23] B. Wang, S. Bao, S. Vinnikova, P.Ghanta, S. Wang, npj Flexible
Electron. 2017, 1, 5.
[24] H.Li, Y.Cao, Z.Wang, X.Feng, Opt. Mater. Express 2019, 9, 4023.
[25] H.Jiang, D.-Y.Khang, J.Song, Y.Sun, Y.Huang, J. A.Rogers, Proc.
Natl. Acad. Sci. USA 2007, 104, 15607.
[26] J. Song, H. Jiang, Z. J. Liu, D. Y.Khang, Y.Huang, J. A. Rogers,
C.Lu, C. G.Koh, Int. J. Solids Struct. 2008, 45, 3107.
[27] S. Wang, J.Song, D.-H. Kim, Y. Huang, J. A. Rogers, Appl. Phys.
Lett. 2008, 93, 023126.
[28] Y.Ma, K.-I.Jang, L. Wang, H. N.Jung, J. W.Kwak, Y.Xue, H.Chen,
Y.Yang, D.Shi, X.Feng, J. A.Rogers, Y.Huang, Adv. Funct. Mater.
2016, 26, 5345.
[29] S. Kim, J. M. Shin, J.-H. Hong, G. S. Park, H. Joo, J. H. Park,
G.Lee, J. Yoon, H.Shin, S.-C.Jo, C.Lee, J.Kwag, Dig. Tech. Pap. -
Soc. Inf. Disp. Int. Symp. 2019, 50, 1194.
Adv. Mater. Technol. 2023, 2201067
2365709x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/admt.202201067 by Seoul National University, Wiley Online Library on [08/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com
© 2023 Wiley-VCH GmbH
2201067 (31 of 37)
www.advmattechnol.de
[30] N.Kim, J.Kim, J.Seo, C.Hong, J.Lee, ACS Appl. Mater. Interfaces
2022, 14, 4344.
[31] S. P. Lacour, S.Wagner, R. J. Narayan, T.Li, Z.Suo, J. Appl. Phys.
2006, 100, 014913.
[32] T. C. Shyu, P. F. Damasceno, P. M. Dodd, A. Lamoureux, L. Xu,
M.Shlian, M. Shtein, S. C.Glotzer, N. A.Kotov, Nat. Mater. 2015,
14, 785.
[33] J. C.Yang, S.Lee, B. S.Ma, J.Kim, M. Song, S. Y.Kim, D. W.Kim,
T.-S.Kim, S.Park, Sci. Adv. 2022, 8, eabn3863.
[34] Y. Zhang, S. Xu, H. Fu, J. Lee, J. Su, K.-C. Hwang, J. A. Rogers,
Y.Huang, Soft Matter 2013, 9, 8062.
[35] R.Xu, Y.Zhang, K.Komvopoulos, Mater. Res. Lett. 2019, 7, 110.
[36] Y. Su, S. Wang, Y.Huang, H.Luan, W. Dong, J. A. Fan, Q.Yang,
J. A.Rogers, Y.Huang, Small 2015, 11, 367.
[37] K.Sim, Y.Li, J.Song, C.Yu, Adv. Mater. Technol. 2019, 4, 1800489.
[38] K. Sim, S. Chen, Z. Li, Z. Rao, J. Liu, Y.Lu, S. Jang, F.Ershad,
J.Chen, J.Xiao, C.Yu, Nat. Electron. 2019, 2, 471.
[39] Q.Hua, J. Sun, H.Liu, R.Bao, R. Yu, J.Zhai, C.Pan, Z. L.Wang,
Nat. Commun. 2018, 9, 244.
[40] M. Jablonski, R. Lucchini, F. Bossuyt, T. Vervust, J. Vanfleteren,
J. W. C.De Vries, P.Vena, M.Gonzalez, Microelectron. Reliab. 2015,
55, 143.
[41] K. Zhang, S. Kong, Y. Li, M. Lu, D. Kong, Lab Chip 2019, 19,
2709.
[42] T.Widlund, S.Yang, Y.-Y.Hsu, N. Lu, Int. J. Solids Struct. 2014, 51,
4026.
[43] S.Yang, E.Ng, N.Lu, Extreme Mech. Lett. 2015, 2, 37.
[44] M. Gonzalez, F. Axisa, M. Vanden Bulcke, D. Brosteaux,
B.Vandevelde, J.Vanfleteren, Microelectron. Reliab. 2008, 48, 825.
[45] A. A.Norhidayah, A. A.Saad, M. F. M.Sharif, F. C.Ani, M. Y. T.Ali,
M. S.Ibrahim, Z.Ahmad, Procedia Eng. 2017, 184, 625.
[46] J.Zhao, H.Hu, W.Fang, Z.Bai, W.Zhang, M.Wu, J. Mater. Chem.
A 2021, 9, 5097.
[47] C. A.Silva, J.Lv, L.Yin, I. Jeerapan, G. Innocenzi, F.Soto, Y.Ha,
J.Wang, Adv. Funct. Mater. 2020, 30, 2002041.
[48] Y.Wang, C.Zhu, R.Pfattner, H.Yan, L.Jin, S.Chen, F.Molina-Lopez,
F. Lissel, J. Liu, N. I. Rabiah, Z. Chen, J. W. Chung,
C. Linder, M. F. Toney, B. Murmann, Z. Bao, Sci. Adv. 2017, 3,
e1602076.
[49] S.Choi, S. I.Han, D.Jung, H. J.Hwang, C.Lim, S.Bae, O. K.Park,
C. M.Tschabrunn, M.Lee, S. Y.Bae, J. W.Yu, J. H.Ryu, S.-W.Lee,
K.Park, P. M. Kang, W. B.Lee, R. Nezafat, T.Hyeon, D.-H. Kim,
Nat. Nanotechnol. 2018, 13, 1048.
[50] S. Choi, J. Park, W. Hyun, J. Kim, J. Kim, Y. B. Lee, C. Song,
H. J.Hwang, J. H.Kim, T.Hyeon, D.-H. Kim, ACS Nano 2015, 9,
6626.
[51] A. M. V.Mohan, N.Kim, Y.Gu, A. J.Bandodkar, J.You, R.Kumar,
J. F. Kurniawan, S. Xu, J. Wang, Adv. Mater. Technol. 2017, 2,
1600284.
[52] K. Li, Y. Shuai, X. Cheng, H. Luan, S. Liu, C. Yang, Z. Xue,
Y.Huang, Y.Zhang, Small 2022, 18, 2107879.
[53] R.Li, M.Li, Y.Su, J.Song, X.Ni, Soft Matter 2013, 9, 8476.
[54] L.Xiao, C.Zhu, W.Xiong, Y.Huang, Z.Yin, Micromachines 2018, 9,
392.
[55] T.Kim, H.Lee, W.Jo, T.Kim, S.Yoo, Adv. Mater. Technol. 2020, 5,
2000494.
[56] S.Gandla, H.Gupta, A. R. Pininti, A. Tewari, D.Gupta, RSC Adv.
2016, 6, 107793.
[57] N.Matsuhisa, M. Kaltenbrunner, T.Yokota, H. Jinno, K.Kuribara,
T.Sekitani, T.Someya, Nat. Commun. 2015, 6, 7461.
[58] D. P. J. Cotton, A. Popel, I. M. Graz, S. P.Lacour, J. Appl. Phys.
2011, 109, 054905.
[59] R. Libanori, R. M. Erb, A. Reiser, H. Le Ferrand, M. J. Süess,
R.Spolenak, A. R.Studart, Nat. Commun. 2012, 3, 1265.
[60] R. Moser, G. Kettlgruber, C. M. Siket, M. Drack, I. M. Graz,
U.Cakmak, Z.Major, M.Kaltenbrunner, S.Bauer, Adv. Sci. 2016, 3,
1500396.
[61] I. M.Graz, D. P. J. Cotton, A.Robinson, S. P. Lacour, Appl. Phys.
Lett. 2011, 98, 124101.
[62] M. S. Lim, M. Nam, S. Choi, Y. Jeon, Y. H. Son, S.-M. Lee,
K. C.Choi, Nano Lett. 2020, 20, 1526.
[63] B.Jang, S. Won, J.Kim, J.Kim, M.Oh, H.Lee, J.Kim, Adv. Funct.
Mater. 2022, 32, 2113299.
[64] Z. Zhang, W. Wang, Y. Jiang, Y.-X. Wang, Y.Wu, J.-C.Lai, S. Niu,
C.Xu, C.-C.Shih, C.Wang, H.Yan, L.Galuska, N.Prine, H.-C.Wu,
D. Zhong, G. Chen, N. Matsuhisa, Y. Zheng, Z. Yu, Y. Wang,
R.Dauskardt, X.Gu, J. B. H.Tok, Z.Bao, Nature 2022, 603, 624.
[65] J. Mun, Y. Ochiai, W. Wang, Y. Zheng, Y.-Q. Zheng, H.-C. Wu,
N. Matsuhisa, T.Higashihara, J. B. H. Tok, Y.Yun, Z. Bao, Nat.
Commun. 2021, 12, 3572.
[66] S.Wang, J.Xu, W.Wang, G.-J. N.Wang, R.Rastak, F.Molina-Lopez,
J. W.Chung, S.Niu, V. R.Feig, J.Lopez, T.Lei, S.-K. Kwon, Y.Kim,
A. M. Foudeh, A. Ehrlich, A. Gasperini, Y. Yun, B. Murmann,
J. B. H.Tok, Z.Bao, Nature 2018, 555, 83.
[67] Y.Zheng, S.Zhang, J. B. H. Tok, Z.Bao, J. Am. Chem. Soc. 2022,
144, 4699.
[68] Q.Zhang, J.Liang, Y.Huang, H.Chen, R.Ma, Mater. Chem. Front.
2019, 3, 1032.
[69] J.Xu, S.Wang, G.-J. N.Wang, C.Zhu, S.Luo, L.Jin, X.Gu, S.Chen,
V. R. Feig, J. W. F. To, S.Rondeau-Gagné, J.Park, B. C. Schroeder,
C. Lu, J. Y. Oh, Y. Wang, Y. Kim, H. Yan, R. Sinclair, D. Zhou,
G.Xue, B. Murmann, C. Linder, W.Cai, J. B.-H. Tok, J. W.Chung,
Z.Bao, Science 2017, 355, 59.
[70] J. Liu, J. Wang, Z. Zhang, F. Molina-Lopez, G.-J. N. Wang,
B. C. Schroeder, X.Yan, Y.Zeng, O. Zhao, H. Tran, T.Lei, Y.Lu,
Y.-X.Wang, J. B. H.Tok, R.Dauskardt, J. W.Chung, Y.Yun, Z.Bao,
Nat. Commun. 2020, 11, 3362.
[71] J.-H.Kim, J.-W.Park, Sci. Adv. 2021, 7, eabd9715.
[72] H.Zhou, S. J.Han, A. K. Harit, D. H. Kim, D. Y. Kim, Y. S. Choi,
H. Kwon, K. Kim, G. Go, H. J. Yun, B. H. Hong, M. C. Suh,
S. Y.Ryu, H. Y.Woo, T.Lee, Adv. Mater. 2022, 34, 2203040.
[73] J.Noh, G.-U. Kim, S. Han, S. J.Oh, Y.Jeon, D.Jeong, S. W.Kim,
T.-S.Kim, B. J.Kim, J.-Y.Lee, ACS Energy Lett. 2021, 6, 2512.
[74] N. Matsuhisa, S. Niu, S. J. K. O’Neill, J. Kang, Y. Ochiai,
T. Katsumata, H.-C. Wu, M. Ashizawa, G.-J. N. Wang, D.Zhong,
X.Wang, X. Gong, R.Ning, H.Gong, I. You, Y.Zheng, Z. Zhang,
J. B. H.Tok, X.Chen, Z.Bao, Nature 2021, 600, 246.
[75] Y. Jiang, Z. Zhang, Y.-X. Wang, D. Li, C.-T. Coen, E. Hwaun,
G.Chen, H.-C.Wu, D.Zhong, S.Niu, W.Wang, A.Saberi, J.-C.Lai,
Y. Wu, Y. Wang, A. A. Trotsyuk, K. Y. Loh, C.-C. Shih, W. Xu,
K.Liang, K.Zhang, Y.Bai, G.Gurusankar, W.Hu, W.Jia, Z. Cheng,
R. H.Dauskardt, G. C. Gurtner, J. B. H.Tok, K.Deisseroth, etal.,
Science 2022, 375, 1411.
[76] S.Wang, J. Y.Oh, J. Xu, H.Tran, Z.Bao, Acc. Chem. Res. 2018, 51,
1033.
[77] Y.Lee, H.Zhou, T.-W.Lee, J. Mater. Chem. C 2018, 6, 3538.
[78] Y. Lee, J. Y. Oh, T. R. Kim, X. Gu, Y.Kim, G. N. Wang, H. Wu,
R. Pfattner, J. W. F. To, T. Katsumata, D. Son, J. Kang,
J. R. Matthews, W.Niu, M. He, R. Sinclair, Y. Cui, J. B. -H. Tok,
T.Lee, Z.Bao, Adv. Mater. 2018, 30, 1704401.
[79] Y. Zheng, Z. Yu, S. Zhang, X. Kong, W. Michaels, W. Wang,
G. Chen, D. Liu, J.-C. Lai, N. Prine, W. Zhang, S. Nikzad,
C. B. Cooper, D. Zhong, J. Mun, Z. Zhang, J. Kang, J. B. H. Tok,
I.McCulloch, J.Qin, X.Gu, Z.Bao, Nat. Commun. 2021, 12, 5701.
[80] B. Lee, H. Cho, K. T.Park, J.-S. Kim, M. Park, H. Kim, Y.Hong,
S.Chung, Nat. Commun. 2020, 11, 5948.
[81] M. Yang, S. W. Kim, S. Zhang, D. Y. Park, C.-W. Lee, Y.-H. Ko,
H.Yang, Y.Xiao, G.Chen, M.Li, J. Mater. Chem. C 2018, 6, 7207.
Adv. Mater. Technol. 2023, 2201067
2365709x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/admt.202201067 by Seoul National University, Wiley Online Library on [08/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com
© 2023 Wiley-VCH GmbH
2201067 (32 of 37)
www.advmattechnol.de
[82] Y. Joo, J. Byun, N. Seong, J. Ha, H. Kim, S. Kim, T. Kim, H. Im,
D.Kim, Y.Hong, Nanoscale 2015, 7, 6208.
[83] H.-S.Liu, B.-C.Pan, G.-S.Liou, Nanoscale 2017, 9, 2633.
[84] Y. Ko, J. Oh, K. T. Park, S. Kim, W. Huh, B. J. Sung, J. A. Lim,
S.-S.Lee, H.Kim, ACS Appl. Mater. Interfaces 2019, 11, 37043.
[85] S. Y. Hong, S. M.Jee, Y. Ko, J.Cho, K. H.Lee, B. Yeom, H.Kim,
J. G.Son, ACS Nano 2022, 16, 2271.
[86] K.-Y.Chun, Y.Oh, J. Rho, J.-H. Ahn, Y.-J.Kim, H. R.Choi, S. Baik,
Nat. Nanotechnol. 2010, 5, 853.
[87] Y. Yang, N. Sun, Z. Wen, P.Cheng, H. Zheng, H. Shao, Y. Xia,
C.Chen, H. Lan, X.Xie, C. Zhou, J.Zhong, X.Sun, S.-T.Lee, ACS
Nano 2018, 12, 2027.
[88] M. D. Bartlett, A. Fassler, N. Kazem, E. J. Markvicka, P.Mandal,
C.Majidi, Adv. Mater. 2016, 28, 3726.
[89] C. Pan, E. J. Markvicka, M. H. Malakooti, J. Yan, L. Hu,
K.Matyjaszewski, C.Majidi, Adv. Mater. 2019, 31, 1900663.
[90] A. F. Silva, H. Paisana, T. Fernandes, J. Góis, A. Serra,
J. F. J. Coelho, A. T.Almeida, C. Majidi, M. Tavakoli, Adv. Mater.
Technol. 2020, 5, 2000343.
[91] F. Krisnadi, L. L. Nguyen, Ankit, J. Ma , M. R. Kulkarni,
N.Mathews, M. D.Dickey, Adv. Mater. 2020, 32, 2001642.
[92] H. Wang, Y. Yao, Z.He, W. Rao, L. Hu, S. Chen, J. Lin, J. Gao,
P. Zhang, X. Sun, X. Wang, Y. Cui, Q. Wang, S. Dong, G. Chen,
J.Liu, Adv. Mater. 2019, 31, 1901337.
[93] S.Park, G. Thangavel, K.Parida, S. Li, P. S.Lee, Adv. Mater. 2019,
31, 1805536.
[94] J. Y.Oh, S.Kim, H.-K.Baik, U.Jeong, Adv. Mater. 2016, 28, 4455.
[95] J.Park, H.Yoon, G. Kim, B. Lee, S.Lee, S.Jeong, T. Kim, J.Seo,
S.Chung, Y.Hong, Adv. Funct. Mater. 2019, 29, 1902412.
[96] J. H. Lee, Y. R.Jeong, G. Lee, S. W. Jin, Y. H. Lee, S. Y. Hong,
H.Park, J. W.Kim, S.-S.Lee, J. S. Ha, ACS Appl. Mater. Interfaces
2018, 10, 28027.
[97] M. Vosgueritchian, D. J.Lipomi, Z. Bao, Adv. Funct. Mater. 2012,
22, 421.
[98] S. P. Lacour, S.Wagner, Z. Huang, Z. Suo, Appl. Phys. Lett. 2003,
82, 2404.
[99] X.Huang, Y. Liu, H.Cheng, W.-J. Shin, J. A. Fan, Z.Liu, C.-J. Lu,
G.-W.Kong, K.Chen, D.Patnaik, S.-H.Lee, S.Hage-Ali, Y.Huang,
J. A.Rogers, Adv. Funct. Mater. 2014, 24, 3846.
[100] S.Biswas, A.Schoeberl, Y.Hao, J.Reiprich, T.Stauden, J. Pezoldt,
H. O.Jacobs, Nat. Commun. 2019, 10, 4909.
[101] K. Li, X. Cheng, F. Zhu, L. Li, Z. Xie, H. Luan, Z. Wang, Z. Ji,
H. Wang, F. Liu, Y. Xue, C. Jiang, X. Feng, L. Li, J. A. Rogers,
Y.Huang, Y.Zhang, Adv. Funct. Mater. 2019, 29, 1806630.
[102] S. Xu, Y. Zhang, J. Cho, J. Lee, X. Huang, L. Jia, J. A. Fan, Y. Su,
J. Su, H. Zhang, H. Cheng, B. Lu, C. Yu, C. Chuang, T. Kim,
T.Song, K.Shigeta, S. Kang, C.Dagdeviren, I. Petrov, P. V. Braun,
Y.Huang, U.Paik, J. A.Rogers, Nat. Commun. 2013, 4, 1543.
[103] H. Cho, Y.Lee, B. Lee, J. Byun, S. Chung, Y. Hong, J. Inf. Disp.
2019, 21, 41.
[104] J.Byun, B.Lee, E. Oh, H. Kim, S. Kim, S.Lee, Y.Hong, Sci. Rep.
2017, 7, 45328.
[105] E. Oh, T.Kim, J. Yoon, S. Lee, D. Kim, B. Lee, J. Byun, H. Cho,
J.Ha, Y.Hong, Adv. Funct. Mater. 2018, 28, 1806014.
[106] Z. Huang, Y. Hao, Y. Li, H. Hu, C. Wang, A. Nomoto, T. Pan,
Y.Gu, Y.Chen, T.Zhang, W.Li, Y.Lei, N.Kim, C.Wang, L.Zhang,
J. W. Ward, A. Maralani, X. Li, M. F.Durstock, A. Pisano, Y.Lin,
S.Xu, Nat. Electron. 2018, 1, 473.
[107] J.Byun, E. Oh, B.Lee, S.Kim, S. Lee, Y. Hong, Adv. Funct. Mater.
2017, 27, 1701912.
[108] Q. Jiang, S. Zhang, J. Jiang, W. Fei, Z. Wu, Adv. Mater. Technol.
2021, 6, 2000966.
[109] K. Park, D.-K. Lee, B.-S. Kim, H. Jeon, N.-E. Lee, D. Whang,
H.-J. Lee, Y. J. Kim, J.-H. Ahn, Adv. Funct. Mater. 2010, 20,
3577.
[110] C. W. Park, J. B. Koo, C.-S. Hwang, H. Park, S. G. Im, S.-Y.Lee,
Appl. Phys. Express 2018, 11, 126501.
[111] G.Cantarella, V.Costanza, A.Ferrero, R.Hopf, C. Vogt, M.Varga,
L. Petti, N. Münzenrieder, L. Büthe, G. Salvatore, A. Claville,
L.Bonanomi, A. Daus, S. Knobelspies, C.Daraio, G.Tröster, Adv.
Funct. Mater. 2018, 28, 1705132.
[112] X.Li, M. M.Hasan, H.-M.Kim, J.Jang, IEEE Trans. Electron Devices
2019, 66, 2971.
[113] M.Kim, D. K.Brown, O.Brand, Nat. Commun. 2020, 11, 1002.
[114] P. L. McEuen, M. S. Fuhrer, Hongkun Park, IEEE Trans. Nano-
technol. 2002, 1, 78.
[115] D. Sun, M. Y. Timmermans, Y. Tian, A. G. Nasibulin,
E. I. Kauppinen, S. Kishimoto, T.Mizutani, Y.Ohno, Nat. Nano-
technol. 2011, 6, 156.
[116] K.Chen, W.Gao, S.Emaminejad, D.Kiriya, H.Ota, H. Y. Y.Nyein,
K.Takei, A.Javey, Adv. Mater. 2016, 28, 4397.
[117] S. H.Chae, W. J.Yu, J. J.Bae, D. L.Duong, D.Perello, H. Y.Jeong,
Q. H. Ta, T. H. Ly, Q. A. Vu, M. Yun, X. Duan, Y. H. Lee, Nat.
Mater. 2013, 12, 403.
[118] A.Chortos, G. I.Koleilat, R.Pfattner, D.Kong, P.Lin, R.Nur, T.Lei,
H.Wang, N.Liu, Y.-C.Lai, M.-G.Kim, J. W.Chung, S.Lee, Z.Bao,
Adv. Mater. 2016, 28, 4441.
[119] L. Cai, S. Zhang, J. Miao, Z. Yu, C. Wang, ACS Nano 2016, 10,
11459.
[120] W. Huang, H. Jiao, Q. Huang, J. Zhang, M. Zhang, Nanoscale
2020, 12, 23546.
[121] J. Liang, L. Li, D. Chen, T.Hajagos, Z. Ren, S.-Y.Chou, W. Hu,
Q.Pei, Nat. Commun. 2015, 6, 7647.
[122] S. Y. Hong, M. S. Kim, H. Park, S. W. Jin, Y. R.Jeong, J. W.Kim,
Y. H. Lee, L. Sun, G. Zi, J. S. Ha, Adv. Funct. Mater. 2019, 29,
1807679.
[123] T.Sekitani, Y.Noguchi, K.Hata, T.Fukushima, T.Aida, T.Someya,
Science 2008, 321, 1468.
[124] C.Müller, S. Gori, D. W.Breiby, J. W.Andreasen, H. D.Chanzy,
R. A. J. Janssen, M. M. Nielsen, C. P. Radano, H. Sirringhaus,
P.Smith, N.Stingelin-Stutzmann, Adv. Funct. Mater. 2007, 17, 2674.
[125] J. Y. Oh, S. Rondeau-Gagné, Y.-C. Chiu, A. Chortos, F. Lissel,
G.-J. N. Wang, B. C. Schroeder, T. Kurosawa, J. Lopez,
T. Katsumata, J. Xu, C. Zhu, X. Gu, W.-G. Bae, Y. Kim, L. Jin,
J. W.Chung, J. B. H.Tok, Z.Bao, Nature 2016, 539, 411.
[126] G.-J. N. Wang, L. Shaw, J. Xu, T. Kurosawa, B. C. Schroeder,
J. Y.Oh, S. J.Benight, Z.Bao, Adv. Funct. Mater. 2016, 26, 7254.
[127] C. Lu, W.-Y. Lee, X. Gu, J. Xu, H.-H. Chou, H. Yan, Y.-C. Chiu,
M. He, J. R. Matthews, W. Niu, J. B.-H. Tok, M. F. Toney,
W.-C.Chen, Z.Bao, Adv. Electron. Mater. 2017, 3, 1600311.
[128] Z. C. Smith, Z. M. Wright, A. M. Arnold, G. Sauvé,
R. D. McCullough, S. A. Sydlik, Adv. Electron. Mater. 2017, 3,
1600316.
[129] J. Mun, G. N. Wang, J. Y. Oh, T.Katsumata, F. L. Lee, J. Kang,
H. Wu, F. Lissel, S. Rondeau-Gagné, J. B. H. Tok, Z. Bao, Adv.
Funct. Mater. 2018, 28, 1804222.
[130] Y.-C. Chiang, H.-C. Wu, H.-F. Wen, C.-C. Hung, C.-W. Hong,
C.-C. Kuo, T.Higashihara, W.-C. Chen, Macromolecules 2019, 52,
4396.
[131] J. I. Scott, X. Xue, M. Wang, R. J. Kline, B. C. Homan,
D.Dougherty, C.Zhou, G.Bazan, B. T.O’Connor, ACS Appl. Mater.
Interfaces 2016, 8, 14037.
[132] J. Mun, J. Kang, Y. Zheng, S. Luo, H. Wu, N. Matsuhisa, J. Xu,
G. N.Wang, Y.Yun, G.Xue, J. B. H.Tok, Z.Bao, Adv. Mater. 2019,
31, 1903912.
[133] M.Shin, J. Y.Oh, K.-E.Byun, Y.-J.Lee, B.Kim, H.-K.Baik, J.-J.Park,
U.Jeong, Adv. Mater. 2015, 27, 1255.
[134] J.Xu, H.-C. Wu, C.Zhu, A.Ehrlich, L.Shaw, M.Nikolka, S.Wang,
F.Molina-Lopez, X.Gu, S.Luo, D.Zhou, Y.-H.Kim, G.-J. N.Wang,
K. Gu, V. R. Feig, S. Chen, Y. Kim, T. Katsumata, Y.-Q. Zheng,
Adv. Mater. Technol. 2023, 2201067
2365709x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/admt.202201067 by Seoul National University, Wiley Online Library on [08/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com
© 2023 Wiley-VCH GmbH
2201067 (33 of 37)
www.advmattechnol.de
H. Yan, J. W.Chung, J. Lopez, B.Murmann, Z. Bao, Nat. Mater.
2019, 18, 594.
[135] W. Wang, S. Wang, R. Rastak, Y. Ochiai, S. Niu, Y. Jiang,
P. K.Arunachala, Y.Zheng, J.Xu, N.Matsuhisa, X.Yan, S.-K.Kwon,
M.Miyakawa, Z. Zhang, R.Ning, A. M. Foudeh, Y.Yun, C.Linder,
J. B. H.Tok, Z.Bao, Nat. Electron. 2021, 4, 143.
[136] Y.-Q. Zheng, Y. Liu, D. Zhong, S. Nikzad, S. Liu, Z. Yu, D. Liu,
H.-C.Wu, C.Zhu, J. Li, H.Tran, J. B. H.Tok, Z.Bao, Science 2021,
373, 88.
[137] S. Nakamura, T. Mukai, M. Senoh, Jpn. J. Appl. Phys. 1991, 30,
L1998.
[138] K. Fujita, A. Shinoda, M. Inai, T. Yamamoto, M. Fujii, D.Lovell,
T.Takebe, K.Kobayashi, J. Cryst. Growth 1993, 127, 50.
[139] S. Park, Y. Xiong, R.-H. Kim, P. Elvikis, M. Meitl, D. Kim, J. Wu,
J. Yoon, C. Yu, Z. Liu, Y. Huang, K. Hwang, P. Ferreira, X. Li,
K.Choquette, J. A.Rogers, Science 2009, 325, 977.
[140] R.-H.Kim, D.-H.Kim, J.Xiao, B. H.Kim, S.-I. Park, B. Panilaitis,
R.Ghaari, J.Yao, M.Li, Z.Liu, V.Malyarchuk, D. G.Kim, A.-P.Le,
R. G. Nuzzo, D. L. Kaplan, F. G.Omenetto, Y. Huang, Z. Kang,
J. A.Rogers, Nat. Mater. 2010, 9, 929.
[141] M. Choi, B. Jang, W. Lee, S. Lee, T. W. Kim, H.-J. Lee, J.-H. Kim,
J.-H.Ahn, Adv. Funct. Mater. 2017, 27, 1606005.
[142] H.Hwang, M.Kong, K. Kim, D. Park, S.Lee, S.Park, H.-J. Song,
U.Jeong, Sci. Adv. 2021, 7, eabh0171.
[143] C. W.Tang, S. A.VanSlyke, Appl. Phys. Lett. 1987, 51, 913.
[144] B.Geroy, P.leRoy, C.Prat, Polym. Int. 2006, 55, 572.
[145] Z. B. Wang, M. G.Helander, J.Qiu, D. P. Puzzo, M. T. Greiner,
Z. M.Hudson, S. Wang, Z. W.Liu, Z. H.Lu, Nat. Photonics 2011,
5, 753.
[146] T.-H. Han, Y. Lee, M.-R. Choi, S.-H.Woo, S.-H.Bae, B. H. Hong,
J.-H.Ahn, T.-W.Lee, Nat. Photonics 2012, 6, 105.
[147] T.-H. Han, S.-H. Jeong, Y. Lee, H.-K. Seo, S.-J. Kwon, M.-H.Park,
T.-W.Lee, J. Inf. Disp. 2015, 16, 71.
[148] M.-H. Park, T.-H. Han, Y.-H.Kim, S.-H. Jeong, Y.Lee, H.-K. Seo,
H.Cho, T.-W.Lee, J. Photonics Energy 2015, 5, 053599.
[149] A. Sugimoto, H. Ochi, S. Fujimura, A. Yoshida, T. Miyadera,
M. Tsuchida, IEEE J. Sel. Top. Quantum Electron. 2004,
10, 107.
[150] D.-Y.Kim, Y. C.Han, H. C.Kim, E. G.Jeong, K. C.Choi, Adv. Funct.
Mater. 2015, 25, 7145.
[151] T. Yokota, P. Zalar, M. Kaltenbrunner, H. Jinno, N. Matsuhisa,
H. Kitanosako, Y.Tachibana, W.Yukita, M. Koizumi, T. Someya,
Sci. Adv. 2016, 2, e1501856.
[152] D.Yin, J. Feng, N.-R.Jiang, R.Ma, Y.-F.Liu, H.-B. Sun, ACS Appl.
Mater. Interfaces 2016, 8, 31166.
[153] S. Jeong, H. Yoon, B. Lee, S. Lee, Y.Hong, Adv. Mater. Technol.
2020, 5, 2000231.
[154] Z.-K. Tan, R. S. Moghaddam, M. L. Lai, P. Docampo, R. Higler,
F. Deschler, M.Price, A.Sadhanala, L. M.Pazos, D. Credgington,
F. Hanusch, T.Bein, H. J.Snaith, R. H.Friend, Nat. Nanotechnol.
2014, 9, 687.
[155] Y.-H.Kim, S.Kim, A. Kakekhani, J. Park, J.Park, Y.-H.Lee, H. Xu,
S. Nagane, R. B. Wexler, D.-H. Kim, S. H. Jo, L. Martínez-Sarti,
P. Tan, A. Sadhanala, G.-S. Park, Y.-W.Kim, B. Hu, H. J. Bolink,
S.Yoo, R. H.Friend, A. M.Rappe, T.-W.Lee, Nat. Photonics 2021,
15, 148.
[156] H.Zhou, J.-W.Park, Thin Solid Films 2016, 619, 281.
[157] Y.Li, S. Chou, P.Huang, C.Xiao, X.Liu, Y.Xie, F.Zhao, Y.Huang,
J. Feng, H. Zhong, H. Sun, Q. Pei, Adv. Mater. 2019, 31,
1807516.
[158] S. G. R.Bade, X. Shan, P. T.Hoang, J.Li, T.Geske, L.Cai, Q.Pei,
C.Wang, Z.Yu, Adv. Mater. 2017, 29, 1607053.
[159] D.-H. Jiang, Y.-C. Liao, C.-J. Cho, L. Veeramuthu, F.-C. Liang,
T.-C.Wang, C.-C.Chueh, T.Satoh, S.-H.Tung, C.-C.Kuo, ACS Appl.
Mater. Interfaces 2020, 12, 14408.
[160] S. Adeniji, O. Oyewole, R. Koech, D. Oyewole, J. Cromwell,
R. Ahmed, O. Oyelade, D. Sanni, K. Orisekeh, A. Bello,
W.Soboyejo, Macromol. Mater. Eng. 2021, 306, 2100435.
[161] H.Zhou, J.Park, Y.Lee, J.Park, J.Kim, J. S.Kim, H.Lee, S. H.Jo,
X.Cai, L.Li, X.Sheng, H. J.Yun, J.Park, J.Sun, T.Lee, Adv. Mater.
2020, 32, 2001989.
[162] C.Keplinger, J. Sun, C. C. Foo, P. Rothemund, G. M. Whitesides,
Z.Suo, Science 2013, 341, 984.
[163] C. Larson, B. Peele, S. Li, S. Robinson, M. Totaro, L. Beccai,
B.Mazzolai, R.Shepherd, Science 2016, 351, 1071.
[164] S. Li, B. N. Peele, C. M. Larson, H. Zhao, R. F.Shepherd, Adv.
Mater. 2016, 28, 9770.
[165] Z. Zhang, L. Cui, X. Shi, X. Tian, D. Wang, C. Gu, E. Chen,
X. Cheng, Y. Xu, Y. Hu, J. Zhang, L. Zhou, H. H. Fong, P. Ma,
G.Jiang, X.Sun, B.Zhang, H.Peng, Adv. Mater. 2018, 30, 1800323.
[166] J. Wang, C. Yan, G.Cai, M.Cui, A. Lee-Sie Eh, P. SeeLee, Adv.
Mater. 2016, 28, 4490.
[167] J.Wang, C.Yan, K. J.Chee, P. S.Lee, Adv. Mater. 2015, 27, 2876.
[168] F.Stauer, K.Tybrandt, Adv. Mater. 2016, 28, 7200.
[169] Y. J. Tan, H. Godaba, G. Chen, S. T. M. Tan, G. Wan, G. Li,
P. M. Lee, Y.Cai, S.Li, R. F.Shepherd, J. S.Ho, B. C. K. Tee, Nat.
Mater. 2020, 19, 182.
[170] D. Son, J. Kang, O. Vardoulis, Y. Kim, N. Matsuhisa, J. Y. Oh,
J. W. To, J.Mun, T.Katsumata, Y.Liu, A. F.McGuire, M. Krason,
F. Molina-Lopez, J. Ham, U. Kraft, Y. Lee, Y.Yun, J. B. H. Tok,
Z.Bao, Nat. Nanotechnol. 2018, 13, 1057.
[171] X.Yan, C.Tian, W.Jiang, W.Xia, X.Wei, J. Soc. Inf. Disp. 2014, 22,
245.
[172] H.Ohmae, Y.Tomita, M.Kasahara, J.Schram, E.Smits, J.vanden
Brand, F.Bossuyt, J.Vanfleteren, J.De Baets, SID Symp. Dig. Tech.
Pap. 2015, 46, 102.
[173] K.Tybrandt, J.Vörös, Small 2016, 12, 180.
[174] J. Woo, H. Lee, C. Yi, J. Lee, C. Won, S. Oh, J. Jekal, C. Kwon,
S.Lee, J.Song, B.Choi, K.Jang, T.Lee, Adv. Funct. Mater. 2020, 30,
1910026.
[175] Y.Lee, B. J. Kim, L.Hu, J.Hong, J.-H.Ahn, Mater. Today 2022, 53,
51.
[176] T.Someya, IEEE Spectrum 2021, 58, 38.
[177] Y. Lee, D. S. Kim, S. W. Jin, H. Lee, Y. R. Jeong, I. You, G. Zi,
J. S.Ha, Chem. Eng. J. 2022, 427, 130858.
[178] X. Shi, Y. Zuo, P.Zhai, J.Shen, Y. Yang, Z.Gao, M. Liao, J. Wu,
J. Wang, X. Xu, Q. Tong, B. Zhang, B. Wang, X. Sun, L. Zhang,
Q.Pei, D.Jin, P.Chen, H.Peng, Nature 2021, 591, 240.
[179] A. K. Tripathi, E. C. P. Smits, J.-L. Van Der Steen, M. Cauwe,
R. Verplancke, K.Myny, J. Maas, R. Kuster, S. Sabik, M. Murata,
Y. Tomita, H. Ohmae, J. Van Den Brand, G. Gelinck, in Int.
Meeting on Information Displays, Daegu, Korea 2015.
[180] S. Y.Hong, Y. H.Lee, H.Park, S. W.Jin, Y. R.Jeong, J. Yun, I.You,
G.Zi, J. S.Ha, Adv. Mater. 2016, 28, 930.
[181] J. Kim, M. Lee, H. J. Shim, R. Ghaari, H. R. Cho, D. Son,
Y. H. Jung, M. Soh, C. Choi, S. Jung, K. Chu, D. Jeon, S.-T.Lee,
J. H.Kim, S. H.Choi, T. Hyeon, D.-H. Kim, Nat. Commun. 2014,
5, 5747.
[182] R. Maiti, L.-C. Gerhardt, Z. S. Lee, R. A. Byers, D. Woods,
J. A. Sanz-Herrera, S. E. Franklin, R. Lewis, S. J. Matcher,
M. J.Carré, J. Mech. Behav. Biomed. Mater. 2016, 62, 556.
[183] T. Daeneke, K. Khoshmanesh, N. Mahmood, I. A. de Castro,
D. Esrafilzadeh, S. J. Barrow, M. D. Dickey, K. Kalantar-zadeh,
Chem. Soc. Rev. 2018, 47, 4073.
[184] H.Zhu, S.Wang, M.Zhang, T.Li, G.Hu, D.Kong, npj Flexible Elec-
tron. 2021, 5, 25.
[185] S.Cheng, Z.Wu, Lab Chip 2012, 12, 2782.
[186] D.Gall, J. Appl. Phys. 2020, 127, 050901.
[187] D.Ji, J.Jang, J. H.Park, D.Kim, Y. S.Rim, D. K. Hwang, Y.-Y.Noh,
J. Inf. Disp. 2021, 22, 1.
Adv. Mater. Technol. 2023, 2201067
2365709x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/admt.202201067 by Seoul National University, Wiley Online Library on [08/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com
© 2023 Wiley-VCH GmbH
2201067 (34 of 37)
www.advmattechnol.de
[188] H. Zhu, E. Shin, A. Liu, D. Ji, Y. Xu, Y. Noh, Adv. Funct. Mater.
2020, 30, 1904588.
[189] J. Park, S. Heo, K. Park, M. H. Song, J.-Y. Kim, G. Kyung,
R. S. Ruo, J.-U. Park, F. Bien, npj Flexible Electron 2017,
1, 9.
[190] D. G.Mackanic, T.-H.Chang, Z.Huang, Y.Cui, Z.Bao, Chem. Soc.
Rev. 2020, 49, 4466.
[191] Y. Dai, H. Hu, M. Wang, J. Xu, S. Wang, Nat. Electron. 2021,
4, 17.
[192] S. Liu, D. S. Shah, R. Kramer-Bottiglio, Nat. Mater. 2021,
20, 851.
[193] W.Song, M.Kong, S.Cho, S.Lee, J.Kwon, H.Bin Son, J. H.Song,
D. Lee, G. Song, S. Lee, S. Jung, S. Park, U. Jeong, Adv. Funct.
Mater. 2020, 30, 2003608.
[194] M. S.Kim, S.Kim, J.Choi, S.Kim, C.Han, Y.Lee, Y.Jung, J.Park,
S.Oh, B.-S.Bae, H.Lim, I.Park, ACS Appl. Mater. Interfaces 2022,
14, 1826.
[195] C.Jiang, F. Rist, H.Wang, J.Wallner, H. Pottmann, Comput. Des.
2022, 143, 103146.
[196] D.Hwang, E. J.Barron, A. B. M. T.Haque, M. D.Bartlett, Sci. Rob.
2022, 7, eabg2171.
[197] Y.Bai, H.Wang, Y. Xue, Y.Pan, J.-T. Kim, X.Ni, T.-L.Liu, Y.Yang,
M.Han, Y.Huang, J. A.Rogers, X.Ni, Nature 2022, 609, 701.
[198] S.Wu, W.Hu, Q.Ze, M.Sitti, R. Zhao, Multifunct. Mater. 2020, 3,
042003.
[199] R. M.Neville, F.Scarpa, A.Pirrera, Sci. Rep. 2016, 6, 31067.
[200] M. Zhang, H. Shahsavan, Y. Guo, A. Pena-Francesch, Y. Zhang,
M.Sitti, Adv. Mater. 2021, 33, 2008605.
Yeongjun Lee is a postdoctoral researcher in the Department of Chemical Engineering at Stanford
University, USA. He received his B.S. in the Department of Materials Science and Engineering
(MSE) from Hanyang University, South Korea in 2012. He received his Ph.D. in the MSE from
Pohang University of Science and Technology (POSTECH), South Korea in 2018. He joined the
MSE at Seoul National University, as a postdoctoral researcher and worked at Organic Material
Lab in Samsung Advanced Institute of Technology as a sta researcher (2019-2021). His research
interests include printed electronics, stretchable electronics, neuromorphic electronics, and
bioelectronics.
Hyeon Cho is a Ph.D. candidate in the Department of Electrical and Computer Engineering,
Seoul National University, Republic of Korea. He received his B.S. degree from the Department
of Electrical and Computer Engineering, Seoul National University in 2017. His current research
interests are stretchable electronic skin and soft energy generators for self-powered wearable
electronics.
Hyungsoo Yoon is a Ph.D. candidate in the Department of Electrical and Computer Engineering,
Seoul National University, Republic of Korea. He received his B.S. degree from the Department
of Electrical and Computer Engineering, Seoul National University in 2016. His current research
interests are methodologies for the integration of microdevices onto stretchable and flexible plat-
forms for conformable microelectronics.
Adv. Mater. Technol. 2023, 2201067
2365709x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/admt.202201067 by Seoul National University, Wiley Online Library on [08/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com
© 2023 Wiley-VCH GmbH
2201067 (35 of 37)
www.advmattechnol.de
Hyunbum Kang is currently senior researcher at Samsung Advanced Institute of Technology,
Samsung Electronics. He obtained his Ph.D. degree in the Department of Chemical and
Biomolecular Engineering at Korea Advanced Institute of Science and Technology, Republic of
Korea. His current interests include stretchable electronic materials and devices.
Hyunjun Yoo is a Ph.D. candidate in the Department of Electrical and Computer Engineering,
Seoul National University, Republic of Korea. He received his B.S. degree from the Department
of Electrical and Computer Engineering, Seoul National University in 2017. His current research
interests are various stress mechanisms of carbon nanotube thin film transistors for soft
electronics.
Huanyu Zhou is currently a postdoctoral researcher in Materials Science and Engineering at
Seoul National University, Korea. He received his B.S. in 2015 and M.S. in 2016 from Yonsei
University, and his Ph.D. from Seoul National University in 2022. His current research is mainly
focused on flexible and stretchable devices based on organic and organic–inorganic hybrid
materials.
Sujin Jeong is a Ph.D. candidate at the Department of Electrical and Computer Engineering,
Seoul National University, Republic of Korea. She received her B.S. degree from the Department
of Electrical and Computer Engineering, Seoul National University in 2017. Her current research
interests are polymer-based optoelectronics and its geometrical engineering for stretchable
electronics.
Adv. Mater. Technol. 2023, 2201067
2365709x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/admt.202201067 by Seoul National University, Wiley Online Library on [08/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com
© 2023 Wiley-VCH GmbH
2201067 (36 of 37)
www.advmattechnol.de
Gae Hwang Lee is currently principal researcher at Samsung Advanced Institute of Technology,
Samsung Electronics. He obtained his Ph.D. degree in the Department of Physics from KAIST,
South Korea in 2009. He worked for the technologies of organic photodiodes/transistors related
to next generation electronics including stacked image sensors, stretchable display. His current
interests include stretchable semiconducting materials and wearable sensors for healthcare
monitoring.
Geonhee Kim is a Ph.D. candidate in the Department of Electrical and Computer Engineering,
Seoul National University, Republic of Korea. He received his B.S. degree from the Department
of Electrical and Computer Engineering, Seoul National University in 2016. His current research
interests are solution-processed deformable transparent electrodes and electrohydrodynamic
printing for high-resolution electronics.
Gyeong-Tak Go is a Ph.D. candidate in the Department of Materials Science and Engineering of
Seoul National University, Republic of Korea. He received his B.S. degree from the Department
of Materials Science and Engineering, Seoul National University in 2018. His research interests
include the stretchable electronics and neuromorphic electronics that exploit artificial synapses.
Jiseok Seo is a Ph.D. candidate in the Department of Electrical and Computer Engineering,
Seoul National University, Republic of Korea. He received his B.S. degree from the Department
of Electrical and Computer Engineering, Seoul National University in 2016. His current research
interests are stretchable electronic skin and solution-processed TFT.
Adv. Mater. Technol. 2023, 2201067
2365709x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/admt.202201067 by Seoul National University, Wiley Online Library on [08/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com
© 2023 Wiley-VCH GmbH
2201067 (37 of 37)
www.advmattechnol.de
Tae-Woo Lee is a professor in the Department of MSE at Seoul National University, South Korea.
He received his Ph.D. in Chemical Engineering from KAIST, South Korea in 2002. He joined Bell
Laboratories, USA, as a postdoctoral researcher and worked at Samsung Advanced Institute
of Technology as research sta (2003–2008). He was an associate professor in MSE at Pohang
University of Science and Technology (POSTECH), South Korea, until August 2016. His research
focuses on printed or soft electronics that use organic and organic-inorganic hybrid materials for
flexible/stretchable displays, solid-state lighting, solar energy conversion devices, and bioinspired
neuromorphic devices.
Yongtaek Hong received B.S. and M.S. in Electronics Engineering, Seoul National University
(SNU), Seoul, Korea, and Ph.D. in EE, Univ. Mich., Ann Arbor, MI, USA. He was a senior
research scientist at Display Science & Technology Center of Eastman Kodak Company,
Rochester, NY, USA (2003~2006). Since 2006, he has worked at ECE, SNU as a professor and
now is Head of SNU Entrepreneurship Center, and Vice-Head of SNU R&DB Foundation in
Business Aairs. He was a visiting professor at Chem. Eng., Stanford University (2012~2013).
His research interests include printed/flexible/stretchable thin-film devices, displays, and sen-
sors for wearable and electronic skin applications.
Youngjun Yun is currently principal researcher at Samsung Advanced Institute of Technology,
Samsung Electronics. He obtained his Ph.D. degree in School of Engineering from Durham
University, UK in 2011. He worked for the technologies of organic/oxide transistors related to
next generation display including reflective color displays, holographic displays, ultra-high-defini-
tion displays, and foldable displays. His current interests include flexible and stretchable displays
and skin-like sensors for healthcare monitoring.
Adv. Mater. Technol. 2023, 2201067
2365709x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/admt.202201067 by Seoul National University, Wiley Online Library on [08/02/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
... Electronic materials and devices for stretchable displays is currently a hot topic. 115 Stretching a conventional light-emitting polymer which has one-dimensional π-conjugated chains is typically accompanied by a decrease in charge transport and hence a decrease in device efficiency. However, blends with polymers that have good chain flexibility offer outstanding potential in the field of optoelectronic devices. ...
... Each time novel transistors or innovative fabrication techniques emerge, numerous applications that were previously inaccessible become achievable. This includes certain advancements, like ultrahigh-definition transparent displays and flexible electronic devices [10][11][12]. At present, light-emitting diodes (LEDs), relying on solid-state semiconductors, have emerged as the next evolution in lighting technology following incandescent and fluorescent lamps. ...
Article
Full-text available
The pursuit of p-type semiconductors has garnered considerable attention in academia and industry. Among the potential candidates, copper iodide (CuI) stands out as a highly promising p-type material due to its conductivity, cost-effectiveness, and low environmental impact. CuI can be employed to create thin films with >80% transparency within the visible range (400–750 nm) and utilizing various low-temperature, scalable deposition techniques. This review summarizes the deposition techniques for CuI as a hole-transport material and their performance in perovskite solar cells, thin-film transistors, and light-emitting diodes using diverse processing methods. The preparation methods of making thin films are divided into two categories: wet and neat methods. The advancements in CuI as a hole-transporting material and interface engineering techniques hold promising implications for the continued development of such devices.
Article
Anisotropic conductive adhesives (ACAs) are the preferred interconnection technology for applications that employ large dies, flexible substrates, and ultra fine‐pitch interconnects. Conventional ACAs require relatively high bonding pressures and temperatures, and ultra fine‐pitch applications challenge the trade‐off between low interconnect resistance and risk of short circuits. This study introduces an ACA‐like interconnection technology that addresses these limitations, allowing for low‐pressure, low‐temperature assembly processes with enhanced particle control at the interconnects. Conductive particles are deposited onto a patterned carrier and subsequently transferred to electrical pads using either non‐conductive film or Ag sintering. The Ag sintering process is performed at a low temperature (140 °C) and low bonding pressure (1 N for a 10 × 10 mm ² chip). The capability of controlling the position and number of conductive particles within individual interconnects is demonstrated. This presents possibilities for achieving ultra‐fine pitch interconnects with negligible risk of short circuits, without compromising electrical resistance. It is demonstrated that interconnect resistance can be tuned by varying the number of conductive particles (achieving a resistance as low as 33 mΩ with 25 particles per interconnect). This approach is applicable to scenarios where the bonding force and temperature must remain low due to the nature of substrate materials.
Article
Skin-mountable electronic materials are being intensively evaluated for use in bio-integrated devices that can mutually interact with the human body. Over the past decade, functional electronic materials inspired by the skin are developed with new functionalities to address the limitations of traditional electronic materials for bio-integrated devices. Herein, the recent progress in skin-mountable functional electronic materials for skin-like electronics is introduced with a focus on five perspectives that entail essential functionalities: stretchability, self-healing ability, biocompatibility, breathability, and biodegradability. All functionalities are advanced with each strategy through rational material designs. The skin-mountable functional materials enable the fabrication of bio-integrated electronic devices, which can lead to new paradigms of electronics combining with the human body.
Article
Full-text available
Stretchable and free‐form displays receive significant attention as they hold immense potential for revolutionizing future display technologies. These displays are designed to conform to irregular surfaces and endure mechanical strains, making them well suited for applications in wearable electronics, biomedical devices, and interactive displays. Traditional light‐emitting devices typically employ brittle inorganic and metallic materials, which are not conducive to stretchability. However, replacing these nonflexible components with flexible/stretchable nanomaterials, soft organic materials, or their composites improves the overall flexibility and stretchability of devices. In this review, the roles and opportunities of nanomaterials, such as thin films, 1D nanofibrous materials, and micro/nanoparticles, are highlighted for enhancing the stretchability and overall performance of various types of light‐emitting devices. By leveraging the unique mechanical and electrical properties of nanomaterials, various efforts emerge to push the boundaries of stretchable display technologies and further realize their full potential for diverse applications.
Article
Flexible/stretchable electronics, which are characterized by their ultrathin design, lightweight structure, and excellent mechanical robustness and conformability, have garnered significant attention due to their unprecedented potential in healthcare, advanced robotics, and human-machine interface technologies. An increasing number of low-dimensional nanostructures with exceptional mechanical, electronic, and/or optical properties are being developed for flexible/stretchable electronics to fulfill the functional and application requirements of information sensing, processing, and interactive loops. Compared to the traditional single-layer format, which has a restricted design space, a monolithic three-dimensional (M3D) integrated device architecture offers greater flexibility and stretchability for electronic devices, achieving a high-level of integration to accommodate the state-of-the-art design targets, such as skin-comfort, miniaturization, and multi-functionality. Low-dimensional nanostructures possess small size, unique characteristics, flexible/elastic adaptability, and effective vertical stacking capability, boosting the advancement of M3D-integrated flexible/stretchable systems. In this review, we provide a summary of the typical low-dimensional nanostructures found in semiconductor, interconnect, and substrate materials, and discuss the design rules of flexible/stretchable devices for intelligent sensing and data processing. Furthermore, artificial sensory systems in 3D integration have been reviewed, highlighting the advancements in flexible/stretchable electronics that are deployed with high-density, energy-efficiency, and multi-functionalities. Finally, we discuss the technical challenges and advanced methodologies involved in the design and optimization of low-dimensional nanostructures, to achieve monolithic 3D-integrated flexible/stretchable multi-sensory systems.
Article
Full-text available
Stretchable electronics have attracted tremendous attention amongst academic and industrial communities due to their prospective applications in personal healthcare, human‐activity monitoring, artificial skins, wearable displays, human‐machine interfaces, etc. Other than mechanical robustness, stable performances under complex strains in these devices that are not for strain sensing are equally important for practical applications. Here, a comprehensive summarization of recent advances in stretchable electronics with strain‐resistive performance is presented. First, detailed overviews of intrinsically strain‐resistive stretchable materials, including conductors, semiconductors, and insulators, are given. Then, systematic representations of advanced structures, including helical, serpentine, meshy, wrinkled, and kirigami‐based structures, for strain‐resistive performance are summarized. Next, stretchable arrays and circuits with strain‐resistive performance, that integrate multiple functionalities and enable complex behaviors, are introduced. This review presents a detailed overview of recent progress in stretchable electronics with strain‐resistive performances and provides a guideline for the future development of stretchable electronics.
Article
Full-text available
Dynamic shape-morphing soft materials systems are ubiquitous in living organisms; they are also of rapidly increasing relevance to emerging technologies in soft machines1–3, flexible electronics4,5 and smart medicines⁶. Soft matter equipped with responsive components can switch between designed shapes or structures, but cannot support the types of dynamic morphing capabilities needed to reproduce natural, continuous processes of interest for many applications7–24. Challenges lie in the development of schemes to reprogram target shapes after fabrication, especially when complexities associated with the operating physics and disturbances from the environment can stop the use of deterministic theoretical models to guide inverse design and control strategies25–30. Here we present a mechanical metasurface constructed from a matrix of filamentary metal traces, driven by reprogrammable, distributed Lorentz forces that follow from the passage of electrical currents in the presence of a static magnetic field. The resulting system demonstrates complex, dynamic morphing capabilities with response times within 0.1 second. Implementing an in situ stereo-imaging feedback strategy with a digitally controlled actuation scheme guided by an optimization algorithm yields surfaces that can follow a self-evolving inverse design to morph into a wide range of three-dimensional target shapes with high precision, including an ability to morph against extrinsic or intrinsic perturbations. These concepts support a data-driven approach to the design of dynamic soft matter, with many unique characteristics.
Article
Full-text available
Intrinsically stretchable organic light‐emitting diodes (ISOLEDs) are becoming essential components of wearable electronics. However, the efficiencies of the ISOLEDs have been highly inferior to their rigid counterparts, which is due to the lack of ideal stretchable electrode materials that can overcome the poor charge injection at one‐dimensional metallic nanowire/organic interfaces. We demonstrate highly‐efficient ISOLEDs that use graphene‐based two‐dimensional‐contact stretchable electrodes (TCSEs) that incorporate a graphene layer on top of embedded metallic nanowires. The graphene layer modifies the work function, promotes charge spreading, and impedes inward diffusion of oxygen and moisture. The work function (WF) of 3.57 eV is achieved by forming a strong interfacial dipole after deposition of a newly‐designed conjugated polyelectrolyte with crown ether and anionic sulfonate groups on TCSE; this is the lowest value ever reported among ISOLEDs, which overcomes the existing problem of very poor electron injection in ISOLEDs. Subsequent pressure‐controlled lamination yielded a highly efficient fluorescent ISOLED with an unprecedently high current efficiency of 20.3 cd/A, which even exceeds that of an otherwise‐identical rigid counterpart. Lastly, a three‐inch five‐by‐five passive matrix ISOLED was demonstrated using convex stretching. This work can provide a rational protocol for designing intrinsically stretchable high‐efficiency optoelectronic devices with favorable interfacial electronic structures. This article is protected by copyright. All rights reserved
Article
Full-text available
Integration of rigid components in soft polymer matrix is considered as the most feasible architecture to enable stretchable electronics. However, a method of suppressing cracks at the interface between soft and rigid materials due to excessive and repetitive deformations of various types remains a formidable challenge. Here, we geometrically engineered Ferris wheel-shaped islands (FWIs) capable of effectively suppressing crack propagation at the interface under various deformation modes (stretching, twisting, poking, and crumpling). The optimized FWIs have notable increased strain at failure and fatigue life compared with conventional circle- and square-shaped islands. Stretchable electronics composed of various rigid components (LED and coin cell) were demonstrated using intrinsically stretchable printed electrodes. Furthermore, electronic skin capable of differentiating various tactile stimuli without interference was demonstrated. Our method enables stretchable electronics that can be used under various geometrical forms with notable enhanced durability, enabling stretchable electronics to withstand potentially harsh conditions of everyday usage.
Article
Full-text available
Skin-attachable sensors, which represent the ultimate form of wearable electronic devices that ensure conformal contact with skin, suffer from motion artifact limitations owing to relative changes in position between the sensor and skin during physical activities. In this study, a polarization-selective structure of a skin-conformable photoplethysmographic (PPG) sensor was developed to decrease the amount of scattered light from the epidermis, which is the main cause of motion artifacts. The motion artifacts were suppressed more than 10-fold in comparison with those of rigid sensors. The developed sensor-with two orthogonal polarizers-facilitated successful PPG signal monitoring during wrist angle movements corresponding to high levels of physical activity, enabling continuous monitoring of daily activities, even while exercising for personal health care.
Article
Full-text available
Intrinsically stretchable bioelectronic devices based on soft and conducting organic materials have been regarded as the ideal interface for seamless and biocompatible integration with the human body. A remaining challenge is to combine high mechanical robustness with good electrical conduction, especially when patterned at small feature sizes. We develop a molecular engineering strategy based on a topological supramolecular network, which allows for the decoupling of competing effects from multiple molecular building blocks to meet complex requirements. We obtained simultaneously high conductivity and crack-onset strain in a physiological environment, with direct photopatternability down to the cellular scale. We further collected stable electromyography signals on soft and malleable octopus and performed localized neuromodulation down to single-nucleus precision for controlling organ-specific activities through the delicate brainstem.
Article
Full-text available
Next-generation light-emitting displays on skin should be soft, stretchable and bright1–7. Previously reported stretchable light-emitting devices were mostly based on inorganic nanomaterials, such as light-emitting capacitors, quantum dots or perovskites6–11. They either require high operating voltage or have limited stretchability and brightness, resolution or robustness under strain. On the other hand, intrinsically stretchable polymer materials hold the promise of good strain tolerance12,13. However, realizing high brightness remains a grand challenge for intrinsically stretchable light-emitting diodes. Here we report a material design strategy and fabrication processes to achieve stretchable all-polymer-based light-emitting diodes with high brightness (about 7,450 candela per square metre), current efficiency (about 5.3 candela per ampere) and stretchability (about 100 per cent strain). We fabricate stretchable all-polymer light-emitting diodes coloured red, green and blue, achieving both on-skin wireless powering and real-time displaying of pulse signals. This work signifies a considerable advancement towards high-performance stretchable displays. A material design strategy and fabrication process is described to produce all-polymer light-emitting diodes with high brightness, current efficiency and good mechanical stability, with applications in skin electronics and human–machine interfaces.
Article
Full-text available
Island‐bridge architectures represent a widely used structural design in stretchable inorganic electronics, where deformable interconnects that form the bridge provide system stretchability, and functional components that reside on the islands undergo negligible deformations. These device systems usually experience a common strain concentration phenomenon, i.e., “island effect”, because of the modulus mismatch between the soft elastomer substrate and its on‐top rigid components. Such an island effect can significantly raise the surrounding local strain, therefore increasing the risk of material failure for the interconnects in the vicinity of the islands. In this work, a systematic study of such an island effect through combined theoretical analysis, numerical simulations and experimental measurements is presented. To relieve the island effect, a buffer layer strategy is proposed as a generic route to enhanced stretchabilities of deformable interconnects. Both experimental and numerical results illustrate the applicability of this strategy to 2D serpentine and 3D helical interconnects, as evidenced by the increased stretchabilities (e.g., by 1.5 times with a simple buffer layer, and 2 times with a ring buffer layer, both for serpentine interconnects). The application of the patterned buffer layer strategy in a stretchable light emitting diodes system suggests promising potentials for uses in other functional device systems.
Article
Full-text available
Mechanical metamaterials possess unusual mechanical properties that cannot be found in nature. Auxetic metamaterials have negative Poisson's ratios and tend to expand in a direction perpendicular to the axial extension direction. When the Poisson's ratio of a display circuit board is forced to be −1 by adopting an auxetic metamaterial, a display can be stretchable without image distortion, and this display is called a meta‐display in this study. The critical obstacles to implementing a stretchable display are large stretchability, high deformation uniformity, and low image distortion. The meta‐display overcomes these obstacles by incorporating micro‐LEDs and a kirigami‐based auxetic circuit board. An auxetic meta‐display with a stretchability of 24.5%, Poisson's ratio of −1, and no image distortion under uniaxial stretching is demonstrated. Finally, the roll transfer process enabled the scaling‐up of a 3‐inch meta‐display attachable to surfaces with non‐zero Gaussian curvatures. This conformity to the non‐zero Gaussian curvature helps realize biomedical applications such as wearable display, phototherapy, and skincare.
Article
Displays that can be reversibly stretched in their geometrical layout are highly important in various applications. Stretchable displays are commonly demonstrated using two-dimensional (2D) geometries with stretchable rubber substrates and elastic interconnects. However, mechanical stretch induces deterioration of the resolution per unit area and blurs the displayed image, consequently lowering the display quality. In this study, we demonstrate a morphable 3D structure inspired by origami/kirigami to produce stretchable displays that can preserve the original image quality by maintaining the display pixel density under stretching. The morphable 3D display consists of a 7 × 7 micro-light-emitting diode (LED) pixel array integrated on a transparent epoxy structural frame, which can be stretched up to 100% without affecting the performance of the display. The functional 3D system can create important unconventional opportunities for optoelectronic devices.