ArticlePDF Available

Integrated multilayer stretchable printed circuit boards paving the way for deformable active matrix

Authors:
  • Meta Reality Labs Research

Abstract and Figures

Conventional rigid electronic systems use a number of metallization layers to route all necessary connections to and from isolated surface mount devices using well-established printed circuit board technology. In contrast, present solutions to prepare stretchable electronic systems are typically confined to a single stretchable metallization layer. Crossovers and vertical interconnect accesses remain challenging; consequently, no reliable stretchable printed circuit board (SPCB) method has established. This article reports an industry compatible SPCB manufacturing method that enables multilayer crossovers and vertical interconnect accesses to interconnect isolated devices within an elastomeric matrix. As a demonstration, a stretchable (260%) active matrix with integrated electronic and optoelectronic surface mount devices is shown that can deform reversibly into various 3D shapes including hemispherical, conical or pyramid.
This content is subject to copyright. Terms and conditions apply.
ARTICLE
Integrated multilayer stretchable printed circuit
boards paving the way for deformable active matrix
Shantonu Biswas1, Andreas Schoeberl2, Yufei Hao3, Johannes Reiprich2, Thomas Stauden2, Joerg Pezoldt2&
Heiko O. Jacobs2*
Conventional rigid electronic systems use a number of metallization layers to route all
necessary connections to and from isolated surface mount devices using well-established
printed circuit board technology. In contrast, present solutions to prepare stretchable elec-
tronic systems are typically conned to a single stretchable metallization layer. Crossovers
and vertical interconnect accesses remain challenging; consequently, no reliable stretchable
printed circuit board (SPCB) method has established. This article reports an industry com-
patible SPCB manufacturing method that enables multilayer crossovers and vertical inter-
connect accesses to interconnect isolated devices within an elastomeric matrix. As a
demonstration, a stretchable (260%) active matrix with integrated electronic and optoe-
lectronic surface mount devices is shown that can deform reversibly into various 3D shapes
including hemispherical, conical or pyramid.
https://doi.org/10.1038/s41467-019-12870-7 OPEN
1California NanoSystems Institute, Elings Hall, Building 266, Mesa Road, University of California, Santa Barbara, CA 93106-6105, USA. 2Fachgebiet
Nanotechnologie, Institut für Mikro- und Nanotechnologien MacroNano®, Technische Universität Ilmenau, Gustav-Kirchhoff-Strasse 1, D-98693 Ilmenau,
Germany. 3School of Mechanical Engineering and Automation, Haidian District, Beijing Institute of Road 37, Beihang University, 100191 Beijing, China.
*email: heiko.jacobs@tu-ilmenau.de
NATURE COMMUNICATIONS | (2019) 10:4909 | https://doi.org/10.1038/s41467-019-12870-7 | www.nature.com/naturecommunications 1
1234567890():,;
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Adramatic increase in research activities has been per-
ceived for last decade to enable mechanically stretchable
and deformable functional electronic devices14. Conse-
quently, a large number of stretchable devices have been realized
demonstrating a wide range of diverse applications that includes
soft robotics5,6, actuators7, electronic eye cameras8,epidermal
electronics9, wearable electronics10,11, metamorphic electronics12,13,
edible electronics14, acoustoelectronics15,16, health monitoring
devices1719,smarttextiles
20 to give a few examples. Most of these
demonstrators often use highly specialized technologies and
unconventional materials that make these technologies more
interesting for research, but less favorable for industrial production
for mass people. Till today, there exist no reliable manufacturing
methods that can be generalized as stretchable printed circuit board
(SPCB) technology21.
Conventional rigid printed circuit boards (PCB) typically
consists of more than one metallization layers to route metal
tracks to interconnect surface mount devices (SMDs) using well-
established manufacturing methods, which is one of the main
reasons behind the paramount success of this technology. On the
other hand, stretchable electronics mostly remains limited to a
single active layer with less complex device integration, which is
primarily due to the lack of reliable manufacturing methods.
Although, there are a few lab prototypes of stretchable devices
demonstrating multilayer electronic systems with different func-
tionalities2224, the materials and methods used to realize such
devices are unconventional and rarely suitable for industrial
production. This technological and materials lacking connes the
complexity of demonstrated stretchable electronic devices25. For
instance, even the simplest functional active matrix requires at
least two metallization layers26.
Additionally, vertical interconnect accesses (VIAs) are required
to interconnect between different active layers in the circuit
boards, which is not well-established in the manufacturing pro-
cess of stretchable electronics. Although, a few approaches have
been reported to realize VIAs in stretchable substrates using
liquid alloys27,28 or solid phase materials22,29, again, the methods
and the materials are incompatible to conventional processing.
Thus, an industry-compatible processing of multilayer metal
tracks and reliable VIAs, which are two important elements to
realize a SPCB technology, remains a primary challenge.
Recently, we demonstrated a single layer SPCB method that
enabled on-hard-carrierfabrication using conventional planar
microfabrication techniques that delays use of elastomeric sub-
strate to the end30. The reported method enabled high tem-
perature processing, high alignment and registration, and allowed
conventional chip assembly methods on a rigid carrier. However,
the previously demonstrated methods used only a single active
layer without complex routing of the metal tracks, which limited
the complexity of the circuit and the device12,13,30. In this article,
we engineered a similar method to realize integrated multilayer
SPCB and demonstrate an alternative development towards the
realization of stretchable electronics with higher integration
density capabilities by introducing stable VIAs through inter-
connection between different metallization layers. The method
used in this article is compatible with conventional micro fabri-
cation processes and uses commercially available pristine SMDs.
Here, we report a SPCB method which replaces the rigid
insulator substrate of conventional PCB with a highly stretchable
silicone elastomer (EcoFlex). Demonstrated manufacturing
method to realize SPCB can be divided in two steps, the rst step
uses hard-carrier fabrication method that is fully compatible with
conventional planar microfabrication techniques, where the sec-
ond step introduces elastomeric substrate and no active fabrica-
tion is required during this. This method has several benets over
other demonstrated methods in the eld of stretchable
electronics22 since it delays the introduction of rubber substrate
which enables high temperature processing, higher registration
and alignment accuracy, and allows to use conventional robotic
or advanced self-assembly of SMD dies31. Additionally, the
method allows testing the device functionality on hard-carrier
which is benecial since it allows for the identication of failure
modes of the circuit and device layer before and after we detach,
bend or stretch the structure. Moreover, like conventional PCB
technology, this method allows direct use of SMD chips. To
realize VIAs in the SPCB, a similar method to the conventional
PCB technology is used here. A highly stretchable (elongated up
to 260% of the original length) multilayered integrated SPCB
design is discussed. To demonstrate the applicability, a fully
addressable LED active matrix has been realized. The integrated
LED display can be deformed to various three-dimensional (3D)
geometrical shapes to morph hemisphere, cone, and pyramid.
Results
On hard carrier fabrication. Figure 1shows design and rst part
of the multilayer SPCB manufacturing method on a hard carrier.
As an example, an active matrix LED display segment is realized.
As mentioned, active matrix LED array requires at least two
metallization layers, VIAs, and the integration of transistors and
LEDs (1a) in an array type fashion. From materials and proces-
sing point of view, several elements are important to achieve this.
Figure 1b schematically presents elements of the rst step of
SPCB manufacturing process on hard-carrier. The depicted
method uses a rigid Si wafer (500 µm thick, MicroChemicals,
Ulm, Germany) as a carrier substrate and forgoing processing are
performed on this rigid substrate. The details of the processing
are added in the methods and supplementary methods (Supple-
mentary Fig. 1).
Release and peeling layer: Even though there are several benets
of using a rigid carrier substrate for the processing, one drawback
of this method is that the nal device has to be detached from
the rigid substrate to be stretchable. A few other demonstrators
use sequential transfer methods to retrieve different parts of the
device such as metal tracks, active components etc. from rigid
carriers to a stretchable substrate. This approach becomes highly
challenging when active elements of the device become fairly
small due to the alignments and registrations, and thus achieving
higher integration density becomes difcult. In the depicted
approach, we realize the complete functional electronics on-hard-
carrier and detach the entire device to a stretchable substrate by a
single step transfer technique using predened two sacricial
layers of poly(methyl-methacrylate) (PMMA) and polyimide (PI).
In this approach spin coated PMMA, (1 µm thick, green in
Fig. 1b) and an eight µm thick layer of PI (blue in Fig. 1b) is used
as a release and a peeling layer, respectively. The peeling layer
supports the buildup of the circuit and enables detachment of the
circuit after the fabrication is completed.
Metal 1: One of the major elements of the stretchable electronic
devices is the conductive interconnects where rigid SMDs are
used, since the interconnects directly contribute to the stretch-
ability of the device. In the current demonstrator, we used 10 µm
thick copper (Cu) tracks patterned as stress-adaptive meander
shaped32. Initially a 50 nm/200 nm thick sputter coated seed layer
of Al/Cu is deposited, which is patterned by photolithography to
electroplate 10 µm thick layer of Cu (Supplementary Fig. 2) to
increase the mechanical robustness of the metal tracks (reddish in
Fig. 1b). Thick metal tracks (>5 µm) were found to be more
robust than previously used thin (<1 µm) metallization layers33.
This metallization layer forms the columns in the addressing
system (Fig. 1c). Moreover, a part of it (50 nm Al) serves as a self-
aligned etch mask in a later plasma etching step, necessary to
ARTICLE NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-019-12870-7
2NATURE COMMUNICATIONS | (2019) 10:4909 | https://doi.org/10.1038/s41467-019-12870-7 | www.nature.com/naturecommunications
Content courtesy of Springer Nature, terms of use apply. Rights reserved
remove the sacricial PI layer and the photo-patternable PI
insulation layer.
Insulation layer: Multilayer electronic systems require electrical
isolation in between different metal tracks to prevent electrical
short-circuits. In the depicted case, the rows and the columns are
separated using a photo-patternable PI layer (HD 4100) layer.
Specically, a spin-coated 20 μm thick photo-patternable PI layer
is used to entirely cover the rst metal tracks. As illustrated in the
gure, the intermediate photo-patternable PI layer (Fig. 1d)
serves as an insulation layer. The optical microscope image
(Fig. 1e, Supplementary Fig. 3) shows a top view of one of the
crossing regions. The bottom metal appears darker, because it is
buried underneath the photo-patternable PI layer. The colored
SEM image (Fig. 1f) shows a tilted view. The photo-patternable PI
layer follows the surface prole and covers the top and sidewalls
of the rst metal tracks. This results in a raised surface, and this
surface topography continues to the second metallization layer,
which produces the visible crossing. The thickness of the
insolation layer is 20 μm. This layer is selectively removed by
plasma etching after encapsulation and detachment using the rst
metal tracks (Al) as a mask. As a result, in the nal device
the isolation layer has a similar geometry as metal 1 and can
be stretched. Electrically we found no short-circuits or leakage
currents between individual rows and columns before or after the
detachment of the device.
VIAs: As mentioned earlier, metal tracks in different layers
require to be interconnected through VIAs and realizing reliable
VIAs in SPCB remained a challenge till today. In the depicted
approach, like the conventional PCB method, we use electro-
plated Cu to realize the VIAs. The photo-patternable PI layer is
also photolithographically patterned to dene openings and
locations of the VIAs, which are subsequently lled with 20 μmof
electrodeposited copper. This second electrodeposition step is
necessary in order to ensure good electrical contact in between
the rst metal tracks and the VIAs, which is difcult to achieve
using thin-lm deposition methods through 20 μm deep holes. A
similar electrodeposition method is also used in conventional
PCB technologies to grow thick VIAs. As shown schematically in
Fig. 1g, the VIAs are grown using metal 1 as a seed layer through
predened openings in the photo-patternable PI layer. A plasma
cleaning process is required prior to electrodeposit the VIAs in
order to remove the residues from photo-patternable PI
(Supplementary Fig. 4) to ensure good electrical and stable
mechanical contact between different metal layers in these
regions.
Metal 2: Second metal tracks are directly deposited on top of
the isolation layer and the VIAs ensuring good electrical contact
between metal 1, VIA, and metal 2. Another 10 µm thick layer of
Cu is used as the second metallization layer (reddish layer). The
colored SEM image (Fig. 1h) shows a tilted view (and optical
1 cm
50 µm
eMicroscope image
50 µm
f SEM image
VIA
Metal 2
Metal 1 10 µm
iCross-section
Metal 1
Metal 2
VIA
h SEM image
50 µm
PP PI
VIA
Metal 2
Metal 1
SMD
Metal tracks
crossing
Si
VIA
PI
PP PI
Vg
V
V+
PMMA
PI
a b
Active matrix layout Layers in SPCB cFunctionality test on hard carrier
d Metal tracks crossing
g
PP PI
isolation
Metal 2
Metal 1
VIA in SPCB
Fig. 1 First steps to realize integrated stretchable printed circuit boards. Schematic layout of an active matrix using LEDs and transistors (a), layers build-up
in the SPCB (b) and on-hard-carrierfunctionality test (c). Metal layers crossing: schematics (d) next to corresponding optical microscope image (e), and
colored SEM image (f); the two metal layers are separated by a 20 µm thick photo-patternable PI layer, the layer is black in the microscope and gray in the
SEM image. VIA: schematics (g) next to a colored tilted SEM image (h) depicting the PI peeling layer (turquoise), bottom metal (yellow) (in parts covered
by photo-patternable PI, (darker turquoise), VIA (reddish), and top metal (reddish) and cross-section (i) of the same. PI polyimide, PMMA poly(methyl-
methacrylate), PP PI photo-patternable PI, SMD surface mount device, VIA vertical interconnect access, SPCB stretchable printed circuit board
NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-019-12870-7 ARTICLE
NATURE COMMUNICATIONS | (2019) 10:4909 | https://doi.org/10.1038/s41467-019-12870-7 | www.nature.com/naturecommunications 3
Content courtesy of Springer Nature, terms of use apply. Rights reserved
microscope image in Supplementary Fig. 5 shows top view) of the
VIA region, revealing the PI peeling layer, metal 1, the photo-
patternable PI layer, VIA, and metal 2, and the cross-sectional
SEM image (Fig. 1i) shows different metal layers in the VIA
region.
Solder bump application: The solder bumps are used to make
mechanical and electrical contact in between the metal tracks and
the active components. A low melting point solder (Indalloy
#117) is used for this purpose because the solder is applied by
a parallel dip-coating process in a liquid solder bath34 using a
10 µm thick photoresist mask to dene the solder bump locations
(see Supplementary Fig. 6). However, a higher melting point
solder could be used as well following other solder printing
methods.
Component assembly: The fabrication process is compatible
with various types of chip attachment and component assembly
methods, including solder bump-based interconnects35,ip-chip
die attachment, robotic pick-and-place or engineered self-assembly
using molten solder31. In the demonstrated case, we use a semi-
automated pick-and-place process to mount the components. This
demonstrator contains a number of lab-fabricated (bare dies,
requiring ip-chip die assembly) eld effect µ-transistors (µ-FET,
see Supplementary Fig. 7) (0.5 × 0.5 × 0.5 mm) and an equal
number of commercially-bought standard surface mount LEDs
(1 × 0.6 × 0.2 mm, 459 nm, Creative LED GMBH, Schaan, Für-
stentum Liechtenstein).
On hard carrier functionality tests: The depicted approach
enables on-hard-carrierfunctionality tests. This is different
from other methods, which build on a soft elastomeric rubber
substrate. This is benecial since it allows for the identication of
failure modes of the circuit and device layers before and after we
detach, bend, or stretch the structure. For example, in the
depicted hard carrier functionality tests (Fig. 1c) all the LEDs
function properly and response to the addressing system.
Introducing elastomeric substrate. In this step, an elastomeric
material is introduced in liquid form which will become the nal
stretchable substrate after curing. A wide range of elastomeric
resin is available including silicone or plastic which make this
method compatible with a large variety of substrate materials with
different mechanical and optical properties. Mainly two proces-
sing steps are involved in this part.
Encapsulation and detachment: To detach the device layer
from the hard carrier, we use a castable 3 mm thick and thermo-
curable (room temperature, 15 h) layer of EcoFlex (Smooth-On,
EcoFlex 00-30) as a stretchable encapsulation layer which is
poured evenly over the entire surface of the fabricated device. To
increase the bond strength between the active layers and the
EcoFlex, a preceding 5 min long O
2
plasma activation step is used.
The process is carried out without any mask, meaning the whole
surface is exposed to the plasma which includes the isolation
layer, metal 2, solder, and the SMDs. In general, the plasma cleans
the surfaces, increases the surface roughness and activates the
polymer surface with increasing OH group.
As shown schematically in Fig. 2a (and Supplementary Fig. 8),
the molding process encapsulates the SMDs, meaning the EcoFlex
under ll the SMDs. It is known that elastomers with high
viscosities will require high local pressures to ll in small gaps.
However, we observed that EcoFlex 00-30 (viscosity, η=3000 cP)
can ll smaller gaps than 10 µm (see Supplementary Fig. 9).
Comparing with Polydimethylsiloxane (PDMS), which is well
known for micro-patterning through soft-lithography, has a
viscosity of 3500 cP. Both of these two polymers are strong
candidates as a stretchable substrate for high dense stretchable
electronics.
The detachment process uses the differential interfacial
adhesion of the stacked layers through an interface that can be
detached. Specically, the PI layer has a low level of adhesion to
the PMMA coated carrier and this interface (PMMA PI)
detached during this step. The detachment process works
particularly well using the introduced PI lm, which forms a
uniform non-stretchable and supporting peeling layer beneath the
circuit. Figure 2b shows an image of the detachment process
while the LEDs are turn on and no damage to the device is
introduced during this process. It should be noted that no solvent
is required during the detachment process which increases the
robustness of the method since some electronic components get
damaged in some solvents.
Removal of unwanted PI: After detachment, the Cu metal
tracks continue to be covered with the PI peeling foil (schematics
in Fig. 2b inset and Supplementary Fig. 10), which needs to be
removed. Moreover, large sections of the intermediate PI foil are
treated as an unwanted layer, because a continuous lm of PI
reduces the stretchability of the system. We used electron
cyclotron resonance (ECR) plasma etching process (40 SCCM
O
2
, 10 SCCM CF
4
, 100 W RF power, 30 min at 0.025 mbar) to
accomplish the removal of these two layers. The interesting part is
the second intermediate photo-patternable PI layer which is used
as an isolation layer between two metal layers. This layer will be
removed everywhere, except for the region that is covered with
metal 1. The rst metal track acts as hard mask during the
plasma etching process which means that the photo-patternable
PI layer will have the same stretchable-meander shape structure
as Metal 1. The image in Fig. 2c shows that all the LEDs are
lighting in the EcoFlex substrate after completing the entire
fabrication process.
Device in rubber matrix. Figure 2df shows images of a single
pixel, crossing of two metal tracks, and VIA, respectively, in the
SPCB in EcoFlex matrix. The active devices remain completely
embedded in EcoFlex (surrounded and under-lled), which
provides protection during nal stretching operations. The close-
up image reveals the bottom side of a single pixel in the active
matrix showing the transistor and the LED (Fig. 2d). The bottom
provides access to the metallization layer, meaning the structure
has a surface mount like geometry. Pads are accessible from the
bottom and all other elements are embedded and surrounded
with silicone; this is different from methods that build on top of
an elastomeric carrier. The SEM image (Fig. 2e) shows the
crossing of two metal tracks in the EcoFlex. After detachment, the
metal 1 becomes on top and metal 2 encapsulates in the EcoFlex
matrix. The insulator (photo-patternable PI layer follows the
shape of the meander of metal 1 and no critical alignment is
required to prevent short-circuits. Figure 2f shows a SEM image
of a VIA intentionally lifted off from the EcoFlex substrate con-
necting two metal tracks.
Reliability of VIAs. The VIAs play a major role in the system
with more than one metallization layers, and are the key
challenges in the eld of multilayer stretchable electronics.
Specically, the area of the VIAs had to be optimized in order
to achieve fully functioning arrays. The results of this optimi-
zation are summarized in Fig. 3with computer simulated stress
prole at the VIA locations while stretched. VIAs connecting
bottom and top metal track in an open location (Fig. 3a, b) and
VIAs connecting a metal track to one of the contact pads of a
component (Fig. 3c, d) can be distinguished. A goal was to
establish the maximum level of uniaxial elongation of the sys-
tem to cause an electrical discontinuity. The measured values of
the elongation ranged up to 260% of the original length can be
ARTICLE NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-019-12870-7
4NATURE COMMUNICATIONS | (2019) 10:4909 | https://doi.org/10.1038/s41467-019-12870-7 | www.nature.com/naturecommunications
Content courtesy of Springer Nature, terms of use apply. Rights reserved
achieved. Considering VIAs to the contact pads, it was found
that maximizing the footprint is benecial. An increase in the
VIA size from 50 × 70 µm to 350 × 500 µm improved the
elongation limit from 135 to 260%. As a comparison, in a
previous report, the elongation limit of a single metal layer
system was 320%32. This is interesting since the shape of the
meander was identical to the one reported here. Clearly, the
reported VIA limits the stretchability at the current state.
Moreover, the location of the VIA within the system has an
effect. For example, VIAs between metal tracks show a different
size dependent failure rate mechanism (Fig. 3a, c). Again, small
(50 × 70 µm) VIAs failed rst.
1 mm 10 µm
Insulator
Si
PMMA
2 cm
bDetachment between
PMMA & PI layer
1 cm
50 µm
VIA
Metal 1
Metal 2 EcoFlex
aOver molding silicone
PI
EcoFlex
Device layer
cActive matrix in EcoFlex
dA single pixel eCrossing in EcoFlex fVIA in EcoFlex
Fig. 2 Second steps to realize integrated stretchable printed circuit boards. Schematic of silicone encapsulation of the device on hard carrier a.A
photograph of the detachment process of the device layer from the hard carrier while the device is under operation (b) and functionality test of the nal
device in EcoFlex (c). Device in rubber matrix: Photograph of a single pixel of the active matrix encapsulated in EcoFlex showing a transistor and a LED (d).
SEM image of a crossing of two metal layers in EcoFlex (e) and a VIA lifted off from the silicone substrate connecting two metal tracks (f)
0.2 mm
Gate
contact
1 cm
2.2 cm
VIA
Stress profile
at 125%
elongation
0.2 mm
0.5 mm
PP PI
115% 165% 190% Maximum level of
elongation 135% 205% 260%
Metal 1
Metal 2
Silicone
Cu
50 × 70 200 × 275 350 × 500 VIA size in µm50 × 70 200 × 275 350 × 500
h
g
d
f
e
b
ac
r = 150 µm
w = 50 µm
t = 10 µm
PP PI = 20 µm
VIA connecting to one of the contact padsVIA connecting two metal tracks in an open location
Computer simulated stress profile while elongated 125% to its original length
Fig. 3 VIA designs affecting the maximum level of elongation. a,bVIAs connecting bottom and top metal tracks in an open location and c,dVIAs
connecting a metal track to one of the contact pads of a SMD. Optical microscope photographs of the VIAs (a,c) and computer simulated stress prole in
the metal tracks while elongated 125% to its original length (b,d). The dimensions of the VIAs and corresponding maximum level of uniaxial elongation are
shown in the table. e,fResults of the stretching tests using a stress adaptive meander shaped metal track in unstretched condition (e) and while elongated
to 220% (up to 260%) to its original length (f). gSchematic dimensions of the meander shaped metal tracks and himage of a failure mode in the SPCB
using current design. VIA vertical interconnect access, PP PI photo-patternable PI
NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-019-12870-7 ARTICLE
NATURE COMMUNICATIONS | (2019) 10:4909 | https://doi.org/10.1038/s41467-019-12870-7 | www.nature.com/naturecommunications 5
Content courtesy of Springer Nature, terms of use apply. Rights reserved
These results are supported by computer aided stress prole
simulations at the VIA locations for different VIA dimensions
and conditions (Fig. 3b, d). As can be seen, the local stresses in
the metal tracks connected to the VIAs vary gradually. At the
same time, the stress concentrations for smaller VIAs are higher
than the larger VIAs, which provides higher mechanical stability
to the larger VIAs (see Supplementary Fig. 11). Also technically,
relatively large VIAs are benecial since this will result uniform
deposition of materials during the electrodeposition process.
However, the effects of different locations of VIAs are not well
understood from the simulated stress prole. This is probably due
to the additional support from the mounted component, which is
not the case for the VIAs between two metal tracks.
Figure 3e, f present photographs of an array of the active
matrix while the device is in unstretched condition and while
being stretched to 220% to its original length using a stress
adaptive meander shape metal track, respectively. The design
parameters of the metal track are presented in the schematic
image (Fig. 3g). Experimentally, we observed that the primary
mechanism involved to fail the device is due to the detachment of
the metal tracks from the stretchable substrate which eventually
results electrical shorts between the metal tracks as shown in
Fig. 3h.
It is worth to mention that the rigid components are embedded
and the second metal tracks are submerged within the elastomeric
matrix, meaning the second metallization tracks are surrounded
by the stretchable substrate from three sides. However, the rst
metallization tracks are only attached with the substrate from one
side via the photo-patternable PI layer (see Fig. 2e). Experimen-
tally, we observed that this metal layer is at risk to be peeled off
from the elastomeric substrate after prolonged and improper use
(Fig. 3h, Supplementary Fig. 12). Full encapsulation using an
additional 20 µm thick spin coated layer of EcoFlex decreases the
risk of detachment.
The depicted active matrix contains 36 pixels, 6 arrays, and 6
columns of LEDs and transistors within a rubber substrate. In the
current array, the pitch of each pixel is 1 cm. However, the spatial
resolution of the matrix is not limited by the manufacturing
method since the approach uses standard microfabrication
techniques to dene the metal tracks and the VIAs. The spatial
resolution is primarily limited by the physical dimensions of the
surface mount components. However, since the approach uses
non-stretchable SMDs as functional elements connected via
meander-shape metal tracks, the total stretchability of the device
depends on the stretchable interconnects and the available
stretchable areas in the device, which means that highly dense
rigid SMDs will reduce the total stretchability of the system.
On the other hand, direct use of commercial chips is benecial
since those chips are highly developed, integrated and miniatur-
ized with a wide range of functionalities, which eliminates the
requirements of developing new electrical components. The
miniaturized commercial components in the stretchable electro-
nics will also advance the sensing and actuation capabilities.
Furthermore, their integration into free form metamorphic
systems will allow to target new applications in robotics, wearable
health care, bioengineering, and in biomedicines.
The structure is robust enough to demonstrate deformation
behavior. In the device shown, the integrated LED matrix is
deformed using different methods to morph from a planar shape
(Fig. 4a) to a concave (Fig. 4b) or convex (Fig. 4c) shapes through
air deation or ination, respectively. Additional deformation
involves 3D guided shape or vacuum forming to form cone
(Fig. 4d) or to a pyramid (Fig. 4e), respectively (see Supplemen-
tary Fig. 13). Since the guiding structure is printed using a 3D
printer, a large variety of 3D shapes can be made; other topologies
have been demonstrated previously12,30. Figure 4f (and Supple-
mentary Fig. 14) shows the addressability of the active LED
arrays, where two sides of the pyramid are addressed. In other
words, 18 addressing lines, 36 VIAs, and 252 electrical
connections to the 72 devices remained intact.
Discussion
In summary, we reported a design and the fabrication of a
stretchable electrical wiring with crossovers and VIAs to isolated
devices within an elastomeric matrix. The process shares some
similarities with current PCB based fabrication concepts. The goal
was to nd a solution to transform these commonly rigid struc-
tures into a stretchable and deformable counterpart. Currently,
only two metallization layers were required from a signal routing
point of view. However, the described method can, in principle,
be scaled to greater numbers of metal layers with local VIAs in
between; a simple solution to prepare global VIAs, i.e., VIAs
1 cm 1 cm
1 cm 1 cm 1 cm
1 cm
bc
a
de f
Fig. 4 Deformable LED active matrix realized using SPCB method. The planar aaddressable LED matrix can be deformed to different shapes without
altering its electrical functionality. The deformation involves deation (b), ination (c), 3D guided shapes (d), and vacuum forming (e). The addressability
of the array remains unaltered after mechanical deformation (f)
ARTICLE NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-019-12870-7
6NATURE COMMUNICATIONS | (2019) 10:4909 | https://doi.org/10.1038/s41467-019-12870-7 | www.nature.com/naturecommunications
Content courtesy of Springer Nature, terms of use apply. Rights reserved
crossing the entire stack is yet to be found. A systematic study led
to optimized dimensions of VIA that could achieve a 260% sys-
tem level elongation. Still the VIA remains the most fragile ele-
ment from a system level failure point of view.
The recent research trends in stretchable electronics clearly
indicate that this emerging technology will not be limited to only
lab-based prototypes, it will pursue enormous attraction in
commercial products as well. Devices which are already demon-
strated in stretchable electronics proof that the technology will
nd many new types of applications and will improve the per-
formance of many existing devices in various manners. Thus, the
limitations remain to be technology related. Stretchable electro-
nics continues to be limited when it comes to a large number of
interconnects or multilayer designs with highly integrated elec-
tronics which cannot be manufactured reliably for long-term
performance using the current state-of-the-art. However, once
fully developed, most electronic system known to mankind could
be stretchable and could morph to take on new interesting form
factors in the future. Many interesting shape adaptive functions
could be demonstrated.
Methods
Fabrication of multilayer integrated stretchable printed circuit boards.A
525 µm thick Si wafer (MicroChemicals, Ulm, Germany) was spin coated with a
1 µm thick layer of PMMA (AR-P 6510, Allresist, Strau sberg, Germany) and with
an 8 µm thick layer of PI (PI 2611, HD Microsystem, Neu-Isenburg, Germany) and
cured in a convection oven at 250 °C for 5 hours under N
2
ow. 50 nm/200 nm
thick layers of Al/Cu was sputter deposited on top of plasma-activated PI layer. The
wafer was then patterned for electroplating by photolithography using a negative
resist (AZ 15NXT, MicroChemicals, Ulm, Germany). A 10 µm thick layer of Cu
was electroplated using Cu 100 electrolyte (NB Technologies, Bremen, Germany).
A current density of ~15 mA/cm2was applied to grow a smooth Cu layer at room
temperature. Unwanted Cu and Al were chemically removed using standard Cu
and Al etchant (MicroChemicals, Ulm, Germany).
A 20 µm thick photo-patternable polyimide (HD 4100, HD Microsystem, Neu-
Isenburg, Germany) was spin coated, baked at 150 °C for 10 min on a hotplate, and
patterned for VIA openings by photolithography. A subsequent descumming
process (5 min, 50 W, 50 SCCM O
2
) was performed to remove any residues from
the patterned photo-patternable PI in the opening. Then, a 20 µm tall VIA of Cu
was electroplated at the openings using the rst metal tracks as the seed layer.
Another 20 nm/200 nm of Ti/Cu was sputter deposited as a second
metallization layer and the wafer was then patterned for electroplating by
photolithography. Another 10 µm thick layer of Cu was electroplated on top of the
Cu seed layer. A chemical etching process was carried out to etch the unwanted
Ti/Cu. The wafer was then patterned for soldering by photolithography using AZ
1518 positive resist and baked at 120 °C for 10 min. The pads were coated by dip
coating in a solder bath (Indalloy #117, MP. 47 °C, Indium Corp., NY).
Assembly of the SMD components. The SMD components were assembled on the
wafer following a standard pick-and-place technique. The wafer was heated from the
back to melt the solder and then the surface mounted components were assembled.
Device tests were performed on the wafer to check the interconnections and the
device performance was compared before and after the detachment process.
Encapsulation and detachment. The silicone EcoFlex (Smooth-On, EcoFlex 00-
30) resin was prepared by mixing Part A and Part B (1:1 volume ratio) and by
degassing the mixture in a desiccator. The liquid resin of silicone was poured on
top of the substrate with assembled components and cured overnight at room
temperature. First, EcoFlex was removed manually from the edges of the wafer as
EcoFlex is strongly bonded to bare Si. Then the EcoFlex layer was peeled using the
sacricial PI layer. As a nal step, the sacricial PI peeling layer was etched in ECR
(40 SCCM O
2
+10 SCCM CF
4
, 100 W RF power, 0.025 mbar) for 30 min.
Code availability
The article does not contain any codes. However, the simulation parameters that were
used to support the ndings of this study are available from the corresponding author
upon reasonable request.
Data availability
The data that support the ndings of this study are available from the corresponding
author upon reasonable request. The source data for all gures are provided with
the paper.
Received: 26 April 2019; Accepted: 26 September 2019;
References
1. Rogers, J. A., Someya, T. & Huang, Y. Materials and mechanics for stretchable
electronics. Science 327, 1603 LP1601607 (2010).
2. Sekitani, T. & Someya, T. Stretchable, large-area organic electronics. Adv.
Mater. (2010). https://doi.org/10.1002/adma.200904054.
3. Xu, F., Lu, W. & Zhu, Y. Controlled 3D buckling of silicon nanowires for
stretchable electronics. ACS Nano 5, 672678 (2011).
4. Wagner, S. & Bauer, S. Materials for stretchable electronics. MRS Bull. (2012).
https://doi.org/10.1557/mrs.2012.37.
5. Majidi, C. Soft robotics: a perspectivecurrent trends and prospects for the
future. Soft Robot. 1,511 (2014).
6. Bauer, S. et al. 25th anniversary article: a soft future: from robots and sensor
skin to energy harvesters. Adv. Mater. 26, 149161 (2014).
7. Biswas, S. & Visell, Y. Emerging material technologies for haptics. Adv. Mater.
Technol. 4, 1900042 (2019).
8. Hwang, S.-W. et al. Materials and fabrication processes for transient and
bioresorbable high-performance electronics. Adv. Funct. Mater. 23,40874093
(2013).
9. Lee, S. Y. et al. Water-resistant exible GaN LED on a liquid crystal polymer
substrate for implantable biomedical applications. Nano Energy 1,145151 (2012).
10. Case, J. et al. Stretchable bioelectronics for medical devices and systems. in
Stretchable Bioelectronics for Medical Devices and Systems (2016). https://doi.
org/10.1007/978-3-319-28694-5.
11. Choi, S., Lee, H., Ghaffari, R., Hyeon, T. & Kim, D.-H. Recent advances in
exible and stretchable bio-electronic devices integrated with nanomaterials.
Adv. Mater. 28, 42034218 (2016).
12. Biswas, S. et al. 3D metamorphic stretchable microphone arrays. Adv. Mater.
Technol. 2, 1700131 (2017).
13. Biswas, S., Reiprich, J., Pezoldt, J., Stauden, T. & Jacobs, H. O. Metamorphic
stretchable touchpad. Adv. Mater. Technol. 4, 1800446 (2019).
14. Hwang, S.-W. et al. Biodegradable elastomers and silicon nanomembranes/
nanoribbons for stretchable, transient electronics, and biosensors. Nano Lett.
15, 28012808 (2015).
15. Jin, S. W. et al. Stretchable loudspeaker using liquid metal microchannel. Sci.
Rep. 5, 11695 (2015).
16. Biswas, S. et al. Metamorphic hemispherical microphone array for three-
dimensional acoustics. Appl. Phys. Lett. 111, 043109 (2017).
17. Amjadi, M., Kyung, K.-U., Park, I. & Sitti, M. Stretchable, skin-mountable,
and wearable strain sensors and their potential applications: a review. Adv.
Funct. Mater. 26, 16781698 (2016).
18. Khan, Y., Ostfeld, A. E., Lochner, C. M., Pierre, A. & Arias, A. C. Monitoring
of vital signs with exible and wearable medical devices. Adv. Mater. 28,
43734395 (2016).
19. Zamarayeva, A. M. et al. Flexible and stretchable power sources for wearable
electronics. Sci. Adv. 3, e1602051 (2017).
20. Cherenack, K. & van Pieterson, L. Smart textiles: challenges and opportunities.
J. Appl. Phys. 112, 091301 (2012).
21. Vaneteren, J. et al. Printed circuit board technology inspired stretchable
circuits. MRS Bull. 37, 254260 (2012).
22. Huang, Z. et al. Three-dimensional integrated stretchable electronics. Nat.
Electron. 1, 473480 (2018).
23. Wang, S. et al. Skin electronics from scalable fabrication of an intrinsically
stretchable transistor array. Nature 555,8388 (2018).
24. Yu, K. J. et al. Bioresorbable silicon electronics for transient spatiotemporal
mapping of electrical activity from the cerebral cortex. Nat. Mater. 15,
782791 (2016).
25. Gonzalez, M. et al. Design of metal interconnects for stretchable electronic
circuits. Microelectron. Reliab. 48, 825832 (2008).
26. Verplancke, R. et al. 49-2: Invited Paper: stretchable passive matrix LED display
with thin-lm based interconnects. SID Symp. Dig. Tech. Pap.47, 664667 (2016).
27. Lim, Y. et al. Biaxially stretchable, integrated array of high performance
microsupercapacitors. ACS Nano 8, 1163911650 (2014).
28. Green Marques, D., Alhais Lopes, P., T. de Almeida, A., Majidi, C. & Tavakoli,
M. Reliable interfaces for EGaIn multi-layer stretchable circuits and
microelectronics. Lab Chip 19, 897906 (2019).
29. Guo, L. & DeWeerth, S. P. High-density stretchable electronics: toward an
integrated multilayer composite. Adv. Mater. 22, 40304033 (2010).
30. Biswas, S. et al. Deformable printed circuit boards that enable metamorphic
electronics. NPG Asia Mater. 8, e336e336 (2016).
31. Biswas, S., Mozafari, M., Stauden, T. & Jacobs, H. Surface tension directed
uidic self-assembly of semiconductor chips across length scales and material
boundaries. Micromachines 7, 54 (2016).
NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-019-12870-7 ARTICLE
NATURE COMMUNICATIONS | (2019) 10:4909 | https://doi.org/10.1038/s41467-019-12870-7 | www.nature.com/naturecommunications 7
Content courtesy of Springer Nature, terms of use apply. Rights reserved
32. Biswas, S. et al. Stress-adaptive meander track for stretchable electronics. Flex.
Print. Electron. 3, 032001 (2018).
33. Park, S.-C. et al. Millimeter thin and rubber-like solid-state lighting modules
fabricated using roll-to-roll uidic self-assembly and lamination. Adv. Mater.
27, 36613668 (2015).
34. Park, S.-C. et al. A rst implementation of an automated reel-to-reel uidic
self-assembly machine. Adv. Mater. 26, 59425949 (2014).
35. Kaltwasser, M. et al. Coreshell transformation-imprinted solder bumps
enabling low-temperature uidic self-assembly and self-alignment of chips
and high melting point interconnects. ACS Appl. Mater. Interfaces 10,
4060840613 (2018).
Acknowledgements
The research received nancial support through grants from German Science Foundation
(JA 1023/3-1, JA1023/8-1, STA556/8-1). S. Biswas would like to thank J. Uziel, I Mar-
quardt, D. Schäfer and B. Hartmann for their help.
Author contributions
S.B. and H.O.J. conceived the idea, S.B. and A.S. designed and performed the experiments.
J.R. and T.S. assisted during the experiments. Y.H. performed the computer simulations.S.
B. and J.P. analyzed the data. All authors contributed writing the manuscript.
Competing interests
The authors declare no competing interests.
Additional information
Supplementary information is available for this paper at https://doi.org/10.1038/s41467-
019-12870-7.
Correspondence and requests for materials should be addressed to H.O.J.
Peer review information Nature Communications thanks Jan Vaneteren and the other,
anonymous, reviewer(s) for their contribution to the peer review of this work. Peer
reviewer reports are available.
Reprints and permission information is available at http://www.nature.com/reprints
Publishers note Springer Nature remains neutral with regard to jurisdictional claims in
published maps and institutional afliations.
Open Access This article is licensed under a Creative Commons
Attribution 4.0 International License, which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as long as you give
appropriate credit to the original author(s) and the source, provide a link to the Creative
Commons license, and indicate if changes were made. The images or other third party
material in this article are included in the articles Creative Commons license, unless
indicated otherwise in a credit line to the material. If material is not included in the
articles Creative Commons license and your intended use is not permitted by statutory
regulation or exceeds the permitted use, you will need to obtain permission directly from
the copyright holder. To view a copy of this license, visit http://creativecommons.org/
licenses/by/4.0/.
© The Author(s) 2019
ARTICLE NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-019-12870-7
8NATURE COMMUNICATIONS | (2019) 10:4909 | https://doi.org/10.1038/s41467-019-12870-7 | www.nature.com/naturecommunications
Content courtesy of Springer Nature, terms of use apply. Rights reserved
1.
2.
3.
4.
5.
6.
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:
use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
use bots or other automated methods to access the content or redirect messages
override any security feature or exclusionary protocol; or
share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at
onlineservice@springernature.com
... Wearable and implantable devices utilize flexible circuit boards for tracking biological movements and tissue structures [1,2]. Previous studies have proposed various methods for fabricating flexible substrates, and the most commonly used methods are thin-film processes based on MEMS processes [3,4] and direct drawing processes using inkjet [5][6][7] or screen printing [8,9]. The fabricated flexible substrate offers high flexibility and allows the addition of elasticity [10]. ...
... For interlayer metal wiring, common methods for connecting Vias include plating [3,14] and using conductive resins [8,15,16] substrate due to the composition of the plating solution, and achieving proper Via filling conditions at a high aspect ratio is problematic. On the other hand, using conductive resin raises concerns about long-term durability [17]. ...
Article
Full-text available
Implantable devices utilize flexible substrates due to their ability to conform to the complex shapes and movements of living organisms. However, these devices have a limit to the available power and necessitate low-resistance wiring to handle the extra power consumed on the substrate side. Moreover, complex wiring technology with multilayer wiring is essential to enhance the functionality of implantable devices. Previous studies primarily employed sputtering and printing methods for the fabrication of flexible substrate wiring. However, the wiring resistance is tens to hundreds Ω, causing large power loss. In this study, we aimed to create flexible, low-resistance multilayer circuit boards with resistance less than 1 Ω for implantable devices. Polydimethylsiloxane was used as the substrate material, while platinum was used as the wiring material. The wiring pattern was formed by laser micromachining using an ultrashort pulse laser, and interlayer connections were achieved using Pt Vias fabricated by microwelding. The multilayer circuit board fabricated low impedance wires of 0.25 Ω or less. Furthermore, the wiring demonstrated excellent insulation between wires, even after a 3-hour exposure test in a simulated biological environment, with no short-circuiting issues. Regarding mechanical properties, no significant changes were observed after subjecting the circuit boards to 500 cycles of repeated bending in a bending test with a 6-mm radius. In conclusion, these results indicate that the flexible multilayer circuit boards are well suited for various implantable devices that require low resistance.
... To design such stretchable MRI coils, unconventional materials and solutions that conform to the anatomy of interest are needed. In particular, copper wires are being replaced by alternative conductive materials, such as liquid metals [4] and conductive threads [5]; typically used printed circuit boards (PCBs) are being replaced by stretchable polymers [6] and textiles [7]. Nonetheless, suitable material selection providing optimized elastic capabilities, sufficient electrical conductivity, and adequate biocompatibility has proven to be quite challenging. ...
Article
Full-text available
Wearable sensors are rapidly gaining influence in the diagnostics, monitoring, and treatment of disease, thereby improving patient outcomes. In this review, we aim to explore how these advances can be applied to magnetic resonance imaging (MRI). We begin by (i) introducing limitations in current flexible/stretchable RF coils and then move to the broader field of flexible sensor technology to identify translatable technologies. To this goal, we discuss (ii) emerging materials currently used for sensor substrates, (iii) stretchable conductive materials, (iv) pairing and matching of conductors with substrates, and (v) implementation of lumped elements such as capacitors. Applicable (vi) fabrication methods are presented, and the review concludes with a brief commentary on (vii) the implementation of the discussed sensor technologies in MRI coil applications. The main takeaway of our research is that a large body of work has led to exciting new sensor innovations allowing for stretchable wearables, but further exploration of materials and manufacturing techniques remains necessary, especially when applied to MRI diagnostics.
... This is higher than the 79-88% range achieved with serpentine-shaped PI interconnects encapsulated in TPU, in work aimed at a similar application 18 . Serpentine tracks on elastomer substrates have demonstrated 260% stretch, but without cyclic testing to assess durability 19 www.nature.com/scientificreports/ on the application, and there is a lack of data quantifying the mechanical strain e-textile parts experience in a garment. 30% strain, for 1000 cycles, has been used in similar work 18,56 . ...
Article
Full-text available
The development of stretchable electronic devices is a critical area of research for wearable electronics, particularly electronic textiles (e-textiles), where electronic devices embedded in clothing need to stretch and bend with the body. While stretchable electronics technologies exist, none have been widely adopted. This work presents a novel and potentially transformative approach to stretchable electronics using a ubiquitous structure: the helix. A strip of flexible circuitry (‘e-strip’) is twisted to form a helical ribbon, transforming it from flexible to stretchable. A stretchable core—in this case rubber cord—supports the structure, preventing damage from buckling. Existing helical electronics have only extended to stretchable interconnects between circuit modules, and individual components such as printed helical transistors. Fully stretchable circuits have, until now, only been produced in planar form: flat circuits, either using curved geometry to enable them to stretch, or using inherently stretchable elastomer substrates. Helical e-strips can bend along multiple axes, and repeatedly stretch between 30 and 50%, depending on core material and diameter. LED and temperature sensing helical e-strips are demonstrated, along with design rules for helical e-strip fabrication. Widely available materials and standard fabrication processes were prioritized to maximize scalability and accessibility.
... In general, D k and D f of PIs decrease with temperature increasing due to the water loss in the films; however, the D k and D f of A2EB-ODPA and ODA-TA2EB can remain relatively stable vales with a small change at 10 GHz (Fig. 6), thereby rendering them well-suited for a broad spectrum of temperature environments. In the realm of flexible printed circuit boards (FPC) 43 , one crucial parameter for dielectric materials is their CTE, which ideally should be equal to or lower than that of copper (17 ppm K −1 ). Notably, the CTE value of A2EB-ODPA (7 ppm K −1 ) is considerably lower than that of ODA-TA2EB (29 ppm K −1 ) 44 . ...
Article
Full-text available
Polyimides have emerged as promising dielectric materials for communication equipment, owing to their excellent thermal stability and processability. Nonetheless, a pressing need remains to reduce the high-frequency dissipation factor (Df) of polyimides. Here, we synthesized various polyimides featuring linear backbone structures, finding that polyimides that incorporate a combination of ester groups and ether bonds exhibit low Df values of 0.0015-0.0024 at 10 GHz. Even in high humidity and temperature conditions they maintain low Df values of <0.005 at 10 GHz. To gain insight into the factors influencing this behavior, we conduct a comprehensive study involving aggregation structures and hygroscopic properties. Our findings highlight the pivotal role of high orientation and crystallinity in determining the high-frequency Df of polyimide films.
Article
Full-text available
Recently, flexible/stretchable micro-scale light-emitting diodes (LEDs), with dimensions significantly smaller than conventional diodes used for illuminations, have emerged for promising applications in areas such as deformable displays, wearable devices for healthcare, etc . For such applications, these devices must have some unusual features that common inorganic LEDs do not intrinsically own, including conformability, biocompatibility, mechanical flexibility, etc . This Perspective focuses on summarizing the most recent progress in developing such flexible emitters based on inorganic semiconductors, followed by reviewing their potential applications. Finally, major challenges and future research directions of deformable micro-scale LEDs are presented.
Article
Hybrid MAPbBr 3 quantum dot cellulose papers are fabricated via a one-step, oleic acid/oleylamine-free vacuum filtration method, and the corresponding photodetectors demonstrate self-powered capability, high flexibility, and exceptional stability.
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
Full-text available
The sense of touch is involved in nearly all human activities, but information technologies for displaying tactile sensory information to the skin are rudimentary when compared to state‐of‐the‐art video and audio displays, or to tactile perceptual capabilities. Realizing tactile displays with good perceptual fidelity will require major advances in engineering, design, and fabrication. Research over several decades has highlighted the difficulties of meeting the required performance benchmarks using conventional devices, processes, and techniques. This has highlighted the important role that will be played by new material technologies that can bridge the electronic and mechanical domains. This must occur at the smallest scales, because of the great perceptual spatial and temporal acuity of the sense of touch. The requirements involved also furnish valuable performance benchmarks against which many emerging material technologies are being evaluated. This article highlights recent research and possibilities enabled through new material technologies, ranging from organic electronic materials, to carbon nanomaterials, and a variety of composites. Emerging material technologies are surveyed for the sense of touch, including sensory considerations and requirements, materials, actuation principles, and design and fabrication methods. A conclusion reflects on the main open challenges and future prospects for research in this area.
Article
Full-text available
Conventional touchpads are an example of a common human–machine interface. They are rigid devices with a limited usability without any form of adaptability and conformity to different morphologies necessary for gaming and virtual reality applications. A metamorphic touchpad is an envisioned conceptual approach of shape changing electronics. To demonstrate this, a multipurpose touchpad able to serve as a flexible, 2D stretchable, and 3D metamorphic device is designed and fabricated. The approach replaces the rigid carrier and rigid interconnects by mechanically stretchable structures. The metamorphic touchpad can be wrapped around 3D shapes or undergo reversible topological changes from a conventional planar to a hemispheric shape.
Article
Full-text available
Stretchable electronics is an emerging technology that creates devices with the ability to conform to nonplanar and dynamic surfaces such as the human body. Current stretchable configurations are constrained to single-layer designs due to limited material processing capabilities in soft electronic systems. Here we report a framework for engineering three-dimensional integrated stretchable electronics by combining strategies in material design and advanced microfabrication. Our three-dimensional devices are built layer by layer through transfer printing pre-designed stretchable circuits on elastomers and creating vertical interconnect accesses using laser ablation and controlled soldering. Our approach enables a higher integration density on stretchable substrates than single-layer approaches and allows new functionalities that would be difficult to implement with conventional single-layer designs. Using this engineering framework, we create a stretchable human–machine interface testbed that is based on a four-layer design and offers eight-channel sensing and Bluetooth data communication capabilities. By combining strategies in material design and advanced microfabrication, three-dimensional integrated stretchable electronic devices can be created, including an eight-channel sensing system with Bluetooth communication capabilities that can be used to extract an array of signals from the human body.
Article
Full-text available
Skin-like electronics that can adhere seamlessly to human skin or within the body are highly desirable for applications such as health monitoring, medical treatment, medical implants and biological studies, and for technologies that include human-machine interfaces, soft robotics and augmented reality. Rendering such electronics soft and stretchable-like human skin-would make them more comfortable to wear, and, through increased contact area, would greatly enhance the fidelity of signals acquired from the skin. Structural engineering of rigid inorganic and organic devices has enabled circuit-level stretchability, but this requires sophisticated fabrication techniques and usually suffers from reduced densities of devices within an array. We reasoned that the desired parameters, such as higher mechanical deformability and robustness, improved skin compatibility and higher device density, could be provided by using intrinsically stretchable polymer materials instead. However, the production of intrinsically stretchable materials and devices is still largely in its infancy: such materials have been reported, but functional, intrinsically stretchable electronics have yet to be demonstrated owing to the lack of a scalable fabrication technology. Here we describe a fabrication process that enables high yield and uniformity from a variety of intrinsically stretchable electronic polymers. We demonstrate an intrinsically stretchable polymer transistor array with an unprecedented device density of 347 transistors per square centimetre. The transistors have an average charge-carrier mobility comparable to that of amorphous silicon, varying only slightly (within one order of magnitude) when subjected to 100 per cent strain for 1,000 cycles, without current-voltage hysteresis. Our transistor arrays thus constitute intrinsically stretchable skin electronics, and include an active matrix for sensory arrays, as well as analogue and digital circuit elements. Our process offers a general platform for incorporating other intrinsically stretchable polymer materials, enabling the fabrication of next-generation stretchable skin electronic devices.
Article
Full-text available
This article describes the realization of a metamorphic microphone array. The array morphs from a concave to a planar and then to a convex shape. The morphing array enables a better (12×) sound source localization when compared to existing static and planar arrangements. To enable the realization, a novel stretchable (elongated up to 320% of the original length) printed circuit board fabrication process is reported. The fabrication process enables high-temperature processing, alignment, and registration. Moreover, conventional and otherwise rigid surface mount devices (SMDs) can be used. The process allows the researcher to increase the number of devices (30 SMDs) and interconnects (60 in total) to the required level. As electrical connections, a polyimide-cladded metal track design is reported; the design sustained 11 050 stretch and release cycles. Moreover, 3D reinforcement frames are reported as an effective measure to shield the SMD components from high levels of stress. The resulting products are millimeter-thin stretchable rubber-embedded and electrically interconnected electronic structures with mechanical rubber-membrane-like properties. Morphology changes involve deflation and inflation of the membrane to bulge inward or outward. Finally, the use of 3D-shaped chaperon is discussed to provide additional shape control.
Article
Full-text available
This article describes the realization of a metamorphic stretchable microphone array, which can be inflated by air to morph from a planar to a hemispherical shape. The array undergoes morphological changes to adjust their receive characteristic. To realize this device, a metamorphic printed circuit board technology (m-PCB) is described. The resulting products are millimeter-thin stretchable silicone embedded and electrically interconnected electronic structures with mechanical properties, which resemble a silicone membrane. The microphone array is used to localize a sound source in a 3D space. The results of the planar orientation (resting shape), and the 3D hemispherical orientation after air inflation are compared. The inflated hemispherical microphone array proofs to be better for 3D acoustic localization and/or beam-forming.
Article
We tackle two well-known problems in the fabrication of stretchable electronics: interfacing soft circuit wiring with silicon chips and fabrication of multi-layer circuits. We demonstrate techniques that allow integration of embedded flexible printed circuit boards (FPCBs) populated with microelectronics into soft circuits composed of liquid metal (LM) interconnects. These methods utilize vertical interconnect accesses (VIAs) that are produced by filling LM alloy into cavities formed by laser ablation. The introduced technique is versatile, easy to perform, clean-room free, and results in reliable multi-layer stretchable hybrid circuits that can withstand over 80% of strain. We characterize the fabrication parameters of such VIAs and demonstrated several applications, including a stretchable touchpad and pressure detection film, and an all-integrated multi-layer electromyography (EMG) circuit patch with five active layers including acquisition electrodes, on-board processing and Bluetooth communication modules.
Article
We demonstrate the realization of core−shell transformation imprinted solder bumps to enable low-temperature chip assembly, while providing a route to high-temperature interconnects through transformation. The reported core−shell solder bump uses a lower melting point BiIn-based shell and a higher melting point Sn core in the initial stage. The bumps enable fluidic self-assembly and self-alignment at relatively low temperatures (60−80 °C). The bumps use the high surface free energy of the liquid shell during the self-assembly to capture freely suspended Si dies inside a heated (80 °C) water bath, leading to well-ordered defect-free chip arrays; the molten liquid shell wets the metal contact (binding site) on the chips and yields self-aligned and electrically connected devices. The solid core provides the anchor point to the substrate. After the completion of the assembly, a short reflow raises the melting point, yielding a solid electrical connection. The low melting point liquid diffuses into the high melting point core. The tuning of the material ratios leads to tailored transformation-imprinted solders with high melting points (160−206 °C) in the final structure.
Article
Stretchable metallic interconnects are commonly designed using a meander-shaped metal track with a uniform width. This article reports on a stress-adaptive metal track design which varies in width to accommodate the produced torque in the metal track during stretching. The stress-adaptive design is inspired by computational and experimental studies of two conventional meander-shape metal tracks identifying a common failure mode; specifically, the propagation of torque leading to twist and mechanical fatigue. The understanding gained led to the stress-adaptive design. The stress-adaptive structure is compared with horseshoe- and U-shaped references and shows improvements in the stress distribution, levels of twist, maximum level of elongation(>320%), and required stretch and release cycles(>6000 at 150% elongation)to cause failure in a long term cycling test.
Book
This book highlights recent advances in soft and stretchable biointegrated electronics. A renowned group of authors address key ideas in the materials, processes, mechanics, and devices of soft and stretchable electronics; the wearable electronics systems; and bioinspired and implantable biomedical electronics. Among the topics discussed are liquid metals, stretchable and flexible energy sources, skin-like devices, in vitro neural recording, and more. Special focus is given to recent advances in extremely soft and stretchable bio-inspired electronics with real-world clinical studies that validate the technology. Foundational theoretical and experimental aspects are also covered in relation to the design and application of these biointegrated electronics systems. This is an ideal book for researchers, engineers, and industry professionals involved in developing healthcare devices, medical tools and related instruments relevant to various clinical practices.