ArticlePDF AvailableLiterature Review

Active Materials for Functional Origami

Wiley
Advanced Materials
Authors:

Abstract and Figures

In recent decades, origami has been explored to aid in the design of engineering structures. These structures span multiple scales and have been demonstrated to be used towards various areas such as aerospace, metamaterial, biomedical, robotics, and architectural applications. Conventionally, origami or deployable structures have been actuated by hands, motors, or pneumatic actuators, which can result in heavy or bulky structures. On the other hand, active materials, which reconfigure in response to external stimulus, eliminate the need for external mechanical loads and bulky actuation systems. Thus, in recent years, active materials incorporated with deployable structures have shown promise for remote actuation of light weight, programmable origami. In this review, active materials such as shape memory polymers and alloys, hydrogels, liquid crystal elastomers, magnetic soft materials, and covalent adaptable network polymers, their actuation mechanisms, as well as how they have been utilized for active origami and where these structures are applicable is discussed. Additionally, the state‐of‐the‐art fabrication methods to construct active origami are highlighted. The existing structural modeling strategies for origami, the constitutive models used to describe active materials, and the largest challenges and future directions for active origami research are summarized. This article is protected by copyright. All rights reserved
Introduction to traditional, engineering, and active origami. a) Schematic of traditional origami crane folding process. b) Different types of common engineering origami such as the Miura–ori pattern and associated metamaterials (Adapted with permission.[⁵] Copyright 2015, National Academy of Sciences); flasher origami pattern and deployable solar array (Reproduced with permission from the authors).[¹⁸] waterbomb base origami pattern and waterbomb tube biomedical stent (Reproduced with permission from the authors).[¹²] c) Representation of common actuation strategies for active origami, including heating of LCE hinges (Adapted with permission.[³³] Copyright 2017, The Royal Society of Chemistry); heating of SMP hinges (Adapted with permission.[³⁵] Copyright 2014, IOP Publishing); heating of hydrogel to expel water at hinges (Adapted with permission.[³⁰] Copyright 2014, IOP Publishing); applied external magnetic field to actuate a magnetized MSM (Adapted with permission.[³⁷] Copyright 2018, Springer Nature). d) Additional common origami mechanisms, including instability‐based origami such as the Kresling pattern (Reproduced with permission.[⁷³] Copyright 2020, National Academy of Sciences); curved‐crease origami (as compared to straight‐crease origami) (Adapted with permission.[⁹⁸] Copyright 2020, the Authors, some rights reserved; exclusive licensee AAAS); pop‐up origami initiated by a stretched kirigami substrate and maintained by an elastic layer of material (Adapted with permission.[¹⁰¹] Copyright 2019, Elsevier).
… 
This content is subject to copyright. Terms and conditions apply.
REVIEW
www.advmat.de
Active Materials for Functional Origami
Sophie Leanza, Shuai Wu, Xiaohao Sun, H. Jerry Qi, and Ruike Renee Zhao*
In recent decades, origami has been explored to aid in the design of
engineering structures. These structures span multiple scales and have been
demonstrated to be used toward various areas such as aerospace,
metamaterial, biomedical, robotics, and architectural applications.
Conventionally, origami or deployable structures have been actuated by
hands, motors, or pneumatic actuators, which can result in heavy or bulky
structures. On the other hand, active materials, which reconfigure in response
to external stimulus, eliminate the need for external mechanical loads and
bulky actuation systems. Thus, in recent years, active materials incorporated
with deployable structures have shown promise for remote actuation of light
weight, programmable origami. In this review, active materials such as shape
memory polymers (SMPs) and alloys (SMAs), hydrogels, liquid crystal
elastomers (LCEs), magnetic soft materials (MSMs), and covalent adaptable
network (CAN) polymers, their actuation mechanisms, as well as how they
have been utilized for active origami and where these structures are
applicable is discussed. Additionally, the state-of-the-art fabrication methods
to construct active origami are highlighted. The existing structural modeling
strategies for origami, the constitutive models used to describe active
materials, and the largest challenges and future directions for active origami
research are summarized.
1. Introduction
The art of paper folding transforms 2D sheets into 3D architec-
tures, with the earliest paper folding being traced to 1st century
China when the papermaking process was invented. Origami, the
Japanese paper folding art, began from the 6th century and grad-
ually developed into more meticulous designs, such as the well-
known paper crane (Figure 1a). Origami has remained popular
over the years as a form of art, while it has also become a promi-
nent scientific topic, also referred to as engineering origami
S. Leanza, S. Wu, R. R. Zhao
Department of Mechanical Engineering
Stanford University
Stanford, CA 94305, USA
E-mail: rrzhao@stanford.edu
X. Sun, H. J. Qi
The George W. Woodruff School of Mechanical Engineering
Georgia Institute of Technology
Atlanta, GA 30332, USA
The ORCID identification number(s) for the author(s) of this article
can be found under https://doi.org/10.1002/adma.202302066
DOI: 10.1002/adma.202302066
(Figure 1b), with great relevance to the
fields of mathematics and mechanics.
Its capability of transforming a 2D sheet
into a complicated 3D shape or reshap-
ing a deployed structure into a compact
folded state attracts significant attention
and development in different engineer-
ing fields and research areas, including
aerospace structures,[1,2 ] metamaterials,[3–7]
robotics,[8,9,10,11 ] biomedical devices,[12,13 ]
reconfigurable electronics,[14–16 ] and
architecture,[5,17 ] encompassing a wide
range of length scales, at which pre-
stresses or actuators are built into the
origami system to facilitate the folding.
As mechanical properties of origami are
tunable by the geometrical parameters
of the origami pattern, structures with
versatile mechanical properties such as
metamaterials can be designed (Figure 1b,
left).[5] Properties of origami such as the
auxetic nature of Miura–ori or variable stiff-
ness during folding have been utilized for
metamaterial applications.[3,5 ] In aerospace
applications, a significant change in size
from the deployed to folded configurations
is necessary for efficient storage and/or
transportation of structures such as solar arrays (Figure 1b,
middle),[1,18,19 ] space habitats,[20] and bellows.[21] Through fold-
ing, origami-based aerospace structures can exhibit such im-
mense changes in size, and origami patterns such as the
flasher, Miura–ori, and Yoshimura have been commonly demon-
strated for this purpose. In addition, origami has been used
for biomedical applications where tools or devices capable
of folding and reconfiguration can aid in or improve cer-
tain procedures, such as deployable origami stents (Figure 1b,
right),[12] origami forceps,[22] and origami-inspired surgical
robots.[23]
While the foldability and shape reconfigurability of origami of-
fers significant reconfiguration of structures, the conventional
methods to actuate the shape change has largely relied on me-
chanical loads provided by hands, motors,[24–26 ] or pneumatic
pumps.[4,11,27–29 ] However, the use of manual loading for engi-
neering applications is often impractical, while motors and pneu-
matically driven systems are often bulky and require complicated
control, especially for high degrees-of-freedom (DOF) actuation.
On the other hand, active materials reconfigure in response to ex-
ternal stimuli, eliminating the need for external mechanical loads
and bulky actuation systems.
In recent decades, active materials, also referred to as stimuli-
responsive or smart materials, provide alternative strategies for
Adv. Mater. 2023, 2302066 © 2023 Wiley-VCH GmbH
2302066 (1 of 35)
www.advancedsciencenews.com www.advmat.de
Figure 1. Introduction to traditional, engineering, and active origami. a) Schematic of traditional origami crane folding process. b) Different types
of common engineering origami such as the Miura–ori pattern and associated metamaterials (Adapted with permission.[5] Copyright 2015, National
Academy of Sciences); flasher origami pattern and deployable solar array (Reproduced with permission from the authors).[18] waterbomb base origami
pattern and waterbomb tube biomedical stent (Reproduced with permission from the authors).[12] c) Representation of common actuation strategies for
active origami, including heating of LCE hinges (Adapted with permission.[33] Copyright 2017, The Royal Society of Chemistry); heating of SMP hinges
(Adapted with permission.[35] Copyright 2014, IOP Publishing); heating of hydrogel to expel water at hinges (Adapted with permission.[30] Copyright
2014, IOP Publishing); applied external magnetic field to actuate a magnetized MSM (Adapted with permission.[37] Copyright 2018, Springer Nature). d)
Additional common origami mechanisms, including instability-based origami such as the Kresling pattern (Reproduced with permission.[73] Copyright
2020, National Academy of Sciences); curved-crease origami (as compared to straight-crease origami) (Adapted with permission.[98] Copyright 2020,
the Authors, some rights reserved; exclusive licensee AAAS); pop-up origami initiated by a stretched kirigami substrate and maintained by an elastic
layer of material (Adapted with permission.[101 ] Copyright 2019, Elsevier).
Adv. Mater. 2023, 2302066 © 2023 Wiley-VCH GmbH
2302066 (2 of 35)
15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202302066 by Stanford University, Wiley Online Library on [12/12/2023]. 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 www.advmat.de
the actuation of reconfigurable systems. These materials and
their composites with functional groups/fillers show response to
external stimuli including light, heat, electric or magnetic fields,
and pH. Utilizing active materials, for instance, shape memory
polymers (SMPs) and alloys (SMAs), hydrogels, liquid crystal
elastomers (LCEs), magnetic soft materials (MSMs), and cova-
lent adaptable network (CAN) polymers, conventional origami
has been advanced to active origami.[1,30–36 ] These origami sys-
tems actuated by stimuli-responsive materials have the advan-
tages of being less bulky, as they do not require motors or
other external actuators, light weight, programmable (and of-
tentimes re-programmable), remotely controllable or tetherless
in many cases, as well as capable of selective or distributed ac-
tuation. These merits can enable applications in remote or bi-
ological spaces where conventionally actuated systems are less
viable. To illustrate how these materials generally operate, sev-
eral examples of active origami are shown in Figure 1c, in
which a box with LCE at its hinges can be folded when heat
is applied, causing the LCE to contract.[33] Similarly, a pyra-
mid structure is actuated by its SMP hinges upon heating,[35] a
hydrogel-based pyramid can fold when water is expelled from
its hinges,[30] and a Miura–ori patterned MSM with embed-
ded magnetic particles and precisely programmed magnetization
aligns its magnetization with an external magnetic field and thus
folds.[37]
Given many excellent recent reviews on stimuli-responsive
materials and their applications,[38–44 ] origami-inspired func-
tional designs,[7,45–48 ] and origami modeling,[49] this review will
focus on summarizing the recent advances in origami systems
that are enabled by active materials. We will qualitatively dis-
cuss the different origami mechanisms, material types and their
actuation mechanisms, and advanced manufacturing of active
origami systems and their applications. Section 1 briefly summa-
rizes the common origami mechanisms and different hinge de-
sign strategies for the structural implementation of origami, fol-
lowed by the actuation mechanisms of active materials and their
applications in Section 2. In Section 3, manufacturing methods
for active origami systems are introduced, including molding,
additive manufacturing, and micro/nano-scale fabrication. Next,
theoretical frameworks including both structural modeling and
material constitutive laws that guide the design of active origami
are briefly reviewed in Section 4. Lastly, in Section 5, we provide
a discussion on the challenges and outlooks for the field of active
origami.
1.1. Origami Mechanisms
In this section, different mechanisms which facilitate reconfig-
uration of origami structures are discussed, starting from clas-
sic origami patterns, and then moving on to discuss origami
based on instability, curved-crease origami, and pop-up origami.
It should be noted that certain types of origami fall under sev-
eral of these categories. Additionally, “kirigami”, or the art of pa-
per folding where cuts are permitted, is inherently very closely
tied to the art of origami. It is common for these two arts to be
discussed in tandem,[7,45,46,50–53 ] especially when discussing real-
world applications of origami which require more sophisticated
structural designs and functional materials.
1.1.1. Classic Origami Patterns
Origami patterns transform between their folded and deployed
states through rigid body rotation of panels, which are connected
by hinges. Conventional origami patterns do not require bend-
ing, twisting, or stretching of panels, for example, Miura–ori,[54]
waterbomb,[55,56 ] Tachi–Miura polyhedron,[57] stacked Miura–
ori patterns,[5,58 ] square-twist,[59] Yos h i m u r a , [60] and Ron Resch
origami.[61,62 ] In real paper-based or polymer-based origami
structures, however, deformation occurs most at the hinges while
panel deformation can also occur. This enlarges the origami
design space for patterns such as Kresling,[63–65 ] flasher,[1,18]
and hypar origami,[66,67 ] and allows for higher DOF shape
morphing.[68]
1.1.2. Origami based on Instability
Instabilities are commonly utilized in the folding and deploy-
ment of origami structures. Kresling origami,[69,70 ] for example,
whose pattern is generated from the twist buckling of thin cylin-
drical shells, can be designed to be monostable or bistable. Bista-
bility or multistability of other patterns such as the hypar[66] or
square-twist[59,71 ] can be reached when allowing for panel bend-
ing. Another way that multi-stable origami structures have been
achieved is via stacking of patterns. For example, the stacking of
Miura–ori sheets yields a structure with multiple stable states.[72]
In a similar way, stacked Kresling units have been demonstrated
intensively as multistable structures (Figure 1d, top),[73,74] loco-
motive robots,[75] and metamaterials.[65] Multistable structures
have the advantages of not requiring constant external force or
stimulus for folding/deployment and can additionally allow for
programmable stiffness of structures[73] or can be used for me-
chanical memory applications.[25,76 ]
Additional examples of reconfigurable structures based on
instability can be found in slender structures. Examples in-
clude buckling of thin strips or patterned sheets to generate
3D architectures,[77–79 ] snap buckling of twisted rods to looped
configurations,[80] and twisting of composite structures with pre-
stressed flanges.[81] A related topic is ring origami,[82–85] or slen-
der ring structures capable of folding via buckling instability
when loads are applied. Those studied recently have been gener-
alized to multiple different ring geometries, follow a self-guided
snap-folding mechanism, and have been demonstrated for fold-
able trusses and devices.[86–90 ]
1.1.3. Curved-Crease Origami
Curved-crease origami was first documented in the 1930s[91]
and then expanded to various patterns by artists, for example,
concentric circles[92] and pleated hyperbolic paraboloid,[66,93] in
which introduced curved creases distinguish them from the
straight creases seen with conventional origami patterns (see
Figure 1d, middle).[98] Elastic deformations in both the hinges
and panels during the folding of curved-crease origami are es-
sential and can lead to complex energy landscapes which are
worth exploring for engineering design. While studies on curved-
crease origami have largely been related to the mathematics of
Adv. Mater. 2023, 2302066 © 2023 Wiley-VCH GmbH
2302066 (3 of 35)
15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202302066 by Stanford University, Wiley Online Library on [12/12/2023]. 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 www.advmat.de
Figure 2. Hinge design and considerations for origami structures. a) Hinges for conventional paper-based origami that are introduced by creasing a
sheet. b) Several methods to generate compliant hinges for origami with finite thickness including reduced thickness at the hinge, cuts introduced at
the hinge, or reduced material stiffness at the hinge. c) Mechanisms for actuation when active materials are introduced to the structure.
describing different curved-crease patterns,[52] recent studies en-
compass engineering applications as well, with topics including
metamaterials,[94] membranes for aerospace structures,[95] adap-
tive shading systems,[96] high-strength performance foldcores,[97]
and robotic grippers with programmable stiffness and snapping
behaviors.[98]
1.1.4. Pop-Up Origami
An additional type of origami, pop-up origami, is characterized
by 2D sheets that can “pop up”/assemble to 3D configurations.
Pop-up origami is most often achieved through the release of pre-
stressed substrates to trigger out-of-plane deformation of thin
sheets to a variety of different 3D states.[77,99,100 ] Other pop-up
origami, however, can be triggered by multistability or plastic-
ity of kirigami patterns paired with elastic materials (Figure 1d,
bottom),[101 ] external mechanical loading of kirigami,[102 ] lam-
ination of layered sheets,[103 ] or ion beam irradiation of thin
metal films.[104 ] While pop-up origami is applicable to many
size scales, it has shown especial promise for the fabrication of
small-scale electronics,[105 ] biomedical devices,[23,106 ] and recon-
figurable robots[107–109 ] with complex geometries.
1.2. Structural Implementation of Origami
For conventional paper-based origami, creases are introduced by
plastically deforming the sheet (Figure 2a). The creases, whose
stiffness is lower than the undeformed panels, serve as the
hinges about which the panels rotate so that different configura-
tions of the origami can be realized. While paper-based origami
has been widely used for artistic purposes and for some addi-
tional small-scale electronic or sensing applications,[110–112 ] many
other practical applications of origami, however, involve mate-
rials with more appreciable thickness. Because of this, there
have been many works to address the structural design of
origami that accounts for the panel and hinge thickness.[113,114,18 ]
Thus, to enable folding of these more broadly encompassed
origami structures, tactics are needed to generate compliant
hinges. As depicted in Figure 2b, reduction in thickness at
the hinge, introduction of cuts at the hinge, or introduction
of a lower-stiffness material at the hinge are all methods that
make folding possible for these origami structures (see Sec-
tion 4.1 for details on structural modeling of origami). Start-
ing from a structure with the compliant hinge design, there
are several hinge actuation mechanisms enabled by the incor-
poration of active materials that result in the desired folding
of an active origami structure (Figure 2c). By either introduc-
ing a bilayer hinge with a contractible material on top, a bi-
layer hinge with an expandable material on the bottom, or an
inactive hinge attached to active panels capable of generating
torques (most common with MSMs), the same general upward
folding of the panels can be achieved once the active mate-
rial is actuated. From here, it can be seen that there is a con-
siderable design space for structural implementations of active
origami.
2. Active Materials for Origami Actuation
Table 1 summarizes the main active materials that will be dis-
cussed in this review, along with the stimulus that leads to the
actuation, the mechanism by which these materials operate, a
broad summary of performance details such as actuation speed
and stiffness of the materials, and the applications that these
active origami systems have been designed for. The details of
these different active origami will be discussed in their respective
sections.
2.1. Shape Memory Polymers (SMPs)
SMPs can be programmed into temporary shapes and recover
their permanent shape upon external stimuli, such as heat, light,
or magnetic field. The shape of SMPs can be changed and
easily recovered, enabling a wide variety of uses in deployable
aerospace structures,[193 ] biomedical applications,[194 ] and flexi-
ble electronics.[195 ] SMPs have the ability to be programmed into
arbitrary shapes and to maintain this temporary shape at ambi-
ent conditions. In addition, for amorphous polymer-based SMPs,
Adv. Mater. 2023, 2302066 © 2023 Wiley-VCH GmbH
2302066 (4 of 35)
15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202302066 by Stanford University, Wiley Online Library on [12/12/2023]. 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 www.advmat.de
Table 1. Summary of active materials, actuation mechanisms, performance details, and applications.
Material Actuation stimulus Actuation Mechanism Performance Details Applications
Shape memory
polymers (SMPs)
Heat[1,35,115–136 ]
Light[137–143 ]
Humidity[144,145 ]
Chemical[141 ]
Release of pre-stress at
hinges[1,35,115–136,138–145 ]
Moderate actuation time
(s–min)
Higher stiffness and
actuation stress (MPa–GPa)
Oftentimes only have
one-way actuation
Biomedical[131,132 ]
Aerospace structures[1,126,128 ]
Energy harvesting[130 ]
Shock absorption[127,129 ]
Hydrogels Water[59,146,147]
Heat[59,147–152 ]
Light[148,153,154 ]
pH[152,155–157 ]
Ionic strength[156,157 ]
Swelling at bilayer
hinge[59,146–149,151,152,154–157 ]
Swelling with
porosity gradient[153 ]
Swelling with
crosslinking
gradient[155 ]
Slow actuation (min–hrs)
Lower stiffness and
actuation stress (kPa)
Reversible deformation
Operate in aqueous
environments
Metamaterials[158,159 ]
Drug delivery robots[160 ]
Cell encapsulation or excision
robots[161–164 ]
Liquid crystal
elastomers (LCEs)
Heat[33,165–173 ]
Humidity[174,175 ]
Light[176,177 ]
Bilayer hinges[33,177 ]
Fiber actuation at
hinges[33,178–180 ]
Twisted nematic or
splayed hinges[171,173 ]
Fast actuation (ms–s)
Moderate stiffness and
actuation stress (kPa–MPa)
Reversible deformation
Biomimetic actuators[165,168 ]
Locomotive robots[168,172,181 ]
Magnetic soft
materials (MSMs)
Magnetic field[31,37,182–189 ] Body torque of
panels[37,182–186,188–192 ]
Fast actuation (ms–s)
Material properties depend
on matrix material
Reversible deformation
Tetherless actuation
Drug delivery[75]
Biomimetic robots[75,183 ]
Robots for object
transportation[31,183,186 ]
Metamaterials[187,189 ]
their elastic modulus can shift by nearly three to four orders of
magnitude before and after their glass transition.[196 ] This high
modulus (oftentimes on the scale of a few GPa) at low tempera-
tures can offer the advantage of load-carrying capabilities. In ad-
dition, the actuation of SMPs typically takes several seconds to
several minutes.[193,197,198 ]
2.1.1. Shape Memory Mechanism
From a mechanistic point of view, SMPs are network polymers
composed of switching segments (molecular chains) and net-
points (chemical or physical crosslinks),[198 ] as indicated by the
blue lines and black dots, respectively, in Figure 3a. These poly-
mers have a transition temperature (Ttrans)suchasaglasstran-
sition (Tg)ormelting(Tm) temperature, above which the mate-
rial will be programmable. A typical shape memory cycle involves
two steps. In the first step (or the programming step), the SMP is
heated to above Ttrans. It becomes rubbery, exhibits an increase in
chain mobility, and can be easily deformed into a programmed
shape. As the deformation is maintained and the temperature is
lowered to below Ttrans, the segments become frozen and main-
tain their deformed chain configurations as well as the deformed
shape of the SMP. This state is commonly referred to as the tem-
porary shape. During the second step (or the recovery step), the
SMP is heated above its Ttrans again, during which the polymer
chains regain mobility and entropy drives the polymer back to
its permanent shape. An example of the process is shown in
Figure 3b,[115] where an SMP box programmed to be temporar-
ily open (Figure 3b-i) gradually returns to its permanent folded
shape (Figure 3b-ii) upon heating. It should be noted that, for
most SMPs, this is a one-way shape memory process: the SMP
cannot shift between two shapes as it is heated and cooled, and
it must be mechanically programmed again to exhibit another
cycle of the shape memory effect. Two-way shape memory, a re-
versible process in which the material can switch back and forth
between two shapes by heating above and cooling below Ttrans is
also achievable, but typically requires special design of polymer
composites or macromolecular structures.[199–201 ] It should also
be mentioned that multi-shape-SMPs[202,203 ] are those which can
switch between more than two shapes and can be obtained by
using multiple transition temperatures and programming steps.
2.1.2. SMP Origami
Depending on the Ttrans of the material, SMP-based ac-
tive origami has the advantage of maintaining its tempo-
rary shape at ambient conditions. One way that SMPs have
been used to achieve active origami is to utilize pre-stretched
sheets.[116,125,133,138–140,143 ] Since these sheets begin as pre-
stretched, they are initially in their “temporary” shapes and, upon
heating, they relax and thus contract. Origami constructed of
these sheets was first demonstrated by Liu et al.,[139 ] where black
ink (serving as hinges) was printed on either side of a polystyrene
(an SMP) sheet. Under IR light, the black ink absorbed light
and transferred heat to the SMP underneath, causing a tem-
perature gradient through the thickness of the film to result in
bending toward the ink-patterned side. Thus, by patterning ink
on either side of the film, unidirectional or bidirectional folding
(Figure 3c) was achieved under the stimulus of unfocused light.
Folding structures such as a box, pyramid, and bidirectional fold-
ing strips were demonstrated with this approach, and similar ap-
proaches were used later on.[138,140,143 ] Tolley et al.[116] developed
Adv. Mater. 2023, 2302066 © 2023 Wiley-VCH GmbH
2302066 (5 of 35)
15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202302066 by Stanford University, Wiley Online Library on [12/12/2023]. 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 www.advmat.de
Figure 3. SMP mechanism, origami, and applications. a) Shape memory effect of an amorphous SMP. b) One-way shape memory process of an SMP
box upon heating (Adapted with permission.[115 ] Copyright 2010, Wiley). c) Self-folding of pre-stretched sheets by local light absorption (Adapted
with permission.[139 ] Copyright 2012, The Royal Society of Chemistry). d) Self-folding of layered structures with prescribed folding angle under uniform
heating (Adapted with permission.[116 ] Copyright 2014, IOP Publishing). e) Sequentially folded and locked box by hinges of different Ttrans under uniform
heating (Adapted with permission.[117 ] Copyright 2015, Springer Nature). f) Hydrogel-SMP composite structures with reversible actuation (Adapted with
permission.[204 ] Copyright 2016, Springer Nature). g) Magnetic SMP gripper with locking capabilities (Adapted with permission.[207 ] Copyright 2020,
Wiley). h) SMP Kresling-based metamaterial with tunable mechanical properties (Adapted with permission.[127 ] Copyright 2020, Elsevier).
one-way SMP origami structures with predictable folding angles
by placing the SMP sheet between two stiff paper layers, with
the hinges located at the gaps between paper layers (Figure 3d-i).
When heated, the sheet contracted, causing bending at the
hinges, which ceased when the panels were in contact with one
another, resulting in a controlled bending angle (Figure 3d-ii).
With this approach, they demonstrated the folding of a polyhe-
dron, Miura–ori (Figure 3d-iii), and various other origami struc-
tures.
While the hinges discussed above exhibit the same general re-
sponse to a homogeneous stimulus, hinges can also be designed
to fold sequentially and enable complex folding routines. Mao
et al.[117 ] used SMPs with different glass transition temperatures
to design hinges and achieved sequentially self-folding and self-
locking structures. They demonstrated a folding box (Figure 3e),
which requires sequential actuation of hinges to avoid interfer-
ence or collision of panels during the folding process. Guided
by finite element analysis (FEA) simulations and a reduced or-
der model to predict collision of panels during folding, SMP
hinges of three different Ttrans were chosen, in which fast-folding
hinges had low Ttrans and slow-folding hinges had high Ttrans,to
enable the sequential folding of a self-locking box (Figure 3e).
This work emphasized the importance of the planning of fold-
ing sequence in complex active origami structures. Other ap-
proaches have also been taken to achieve sequential folding. For
example, by using the same material as that in their initial[139 ]
work, Liu et al.[140 ] later achieved sequential folding by using
hinges of different colored inks that absorbed IR light with dif-
ferent efficiency, allowing for sequential folding of a variety of
structures.
Adv. Mater. 2023, 2302066 © 2023 Wiley-VCH GmbH
2302066 (6 of 35)
15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202302066 by Stanford University, Wiley Online Library on [12/12/2023]. 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 www.advmat.de
As discussed previously, most SMPs cannot achieve reversible
actuation. Therefore, active origami that solely uses SMPs can-
not typically have desirable reversible folding and unfolding.
One way to realize reversibly foldable active origami is through
SMP composites, in which another active material drives (or pro-
grams) the shape change while the SMP locks it. Mao et al.[204 ]
demonstrated this concept with structures composed of a hydro-
gel, an elastomer, and an SMP. As seen in Figure 3f-i, a struc-
ture was designed with an SMP top layer, elastomer side and bot-
tom layers, and hydrogels confined by columns between the top
and bottom layers. Upon swelling in water and heating to soften
the SMP layer, the structure bent toward the elastomer layer. The
structure was then cooled to stiffen the SMP and thus lock the
temporary shape. After drying of the hydrogel, the locked shape
was still maintained. The original shape of the structure could
be recovered by heating it again, allowing the SMP to return to
its equilibrium shape. This process could be repeated, enabling
reversibly foldable and un-foldable SMP composite structures
(Figure 3f-ii). Recently, this approach was further developed by
simply using an SMP/hydrogel bilayer to enable reversible shape
change by Yuan et al.[205]
Magnetic particles have also been used to establish similar
functionality in SMPs. This type of material capable of reversible
shape transformation has been termed as magnetic SMP (M-
SMP). Photothermally heated M-SMP and inductively heated M-
SMP have been developed by Liu et al.[206 ] and Ze et al.,[207 ] re-
spectively. In the work by Ze et al.,[207 ] an M-SMP with two differ-
ent types of magnetic particles (NdFeB and Fe3O4) was created.
The Fe3O4particles enabled inductive heating of the SMP under
a high frequency alternating current (AC) magnetic field (heating
magnetic field), while magnetized NdFeB particles drove shape
change of the SMP by aligning the polymer’s magnetization di-
rection to an actuation magnetic field. When the heating mag-
netic field was turned on, the M-SMP became soft and could
easily be deformed into a desirable shape by the actuation mag-
netic field. Further, by keeping the actuation magnetic field on
but ceasing the heating magnetic field, the M-SMP cooled down
until it stiffened. After the actuation magnetic field was turned
off, the temporary shape was fixed, and this method was demon-
strated for the use of a soft gripper (Figure 3g). The incorporation
of SMP with other active materials such as LCEs has also been
demonstrated recently,[208] and it can be seen that SMP compos-
ites allow for various interesting deformations and functionali-
ties.
SMPs enable a wide variety of applications, across many differ-
ent fields of research. One application of SMPs has been active
metamaterials,[119,125,129,209 ] which can exhibit special mechanical
properties. For example, Tao et al.[127] designed SMP Kresling-
based metamaterials in which temperature could be used to con-
trol the load bearing and energy absorption capabilities of the
structures. As shown in Figure 3h, structures could be designed
to fold in sequential ways due to the geometry of the Kresling
origami used, which could be utilized for impact mitigation or
soft robotic applications. Recently, biocompatible shape mem-
ory materials such as keratin[210 ] have also been explored for ac-
tive origami, potentially for use as smart textiles. Chen et al.[1]
demonstrated the use of SMPs in facilitating the deployment
of flasher origami solar arrays for potential aerospace applica-
tions. Additionally, SMPs have been used for other exciting ap-
plications such as robotic joints with tunable stiffness,[211,212 ]
deployable origami sandwich structures,[126 ] origami reflector
antennas,[128 ] carbon nanotube (CNT)-SMP-based foldable solar
evaporators,[130 ] and origami biomedical devices.[131,132 ]
2.2. Hydrogels
Hydrogels are network polymers that can absorb large amounts
of water. Some hydrogels can exhibit macroscopic deformation by
absorbing (swelling) or expelling (de-swelling) water in response
to changes in stimulus such as heat, pH change, ionic strength,
and light. The stiffness and actuation stress of hydrogels typically
lie between a few kPa to several hundred kPa.[213,214] Hydrogel ac-
tuation is slow, as it is driven by a diffusion process. This results
in hydrogel actuation occurring anywhere from tens of minutes
to several hours. Storage of elastic potential energy,[215] utiliza-
tion of instabilities,[216 ] incorporation of particles for hydrogel
composites,[148,153 ] or confined swelling environment for reten-
tion of high osmotic pressures[217 ] have all been taken advantage
of for increasing either the actuation speed, force, or mechanical
strength of hydrogels. Due to their low elastic modulus compa-
rable to biological materials and tissue,[218 ] hydrogels are often
designed by using biocompatible polymers, making them widely
used for biomedical applications, such as patches[219 ] and drug
delivery.[160] Hydrogels are also used in soft robotics including
biomimicking systems,[220 ] locomotive robots,[221–223 ] as well as
active origami.[147,148,156 ]
2.2.1. Hydrogel Mechanism
In general, hydrogel swelling or de-swelling is due to the inter-
action between the polymer chains and the surrounding solvent
molecules. In a good solvent, in which the interaction between
the polymer chains and the solvent is preferred, polymer chains
are arranged as hydrophilic “coils” and the hydrogel can absorb
solvent (such as water). On the other hand, in a poor solvent,
in which the interaction between polymer chains with one an-
other is preferred, polymer chains become hydrophobic “glob-
ules” and the solvent will be expelled. Under different conditions
such as varied temperatures or pH, a solvent can be good or poor
with respect to the polymer chains, leading to a volume phase
transition of the hydrogel.[224,225 ] For example, some hydrogels,
such as poly(N-isopropylacrylamide) (pNIPAM), will repel water
once heated above the hydrogel’s lower critical solution temper-
ature (LCST) and the hydrophilic coils will become hydropho-
bic globules (Figure 4a). It is important to note that swelling of
neat hydrogels is isotropic, so to achieve more complex defor-
mation necessary for origami, tactics such as varied crosslinking
densities,[150,226 ] porosity gradients,[153 ] or structures such as bi-
layers of materials with different swelling capabilities,[147,161,227 ]
are often used. Among these tactics, bilayers are the overwhelm-
ingly popular method for the creation of hydrogel hinges for
origami.
2.2.2. Hydrogel Origami
Bilayer-based hinge design typically consists of a layer of ac-
tive material and a layer of inactive material, or two active
Adv. Mater. 2023, 2302066 © 2023 Wiley-VCH GmbH
2302066 (7 of 35)
15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202302066 by Stanford University, Wiley Online Library on [12/12/2023]. 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 www.advmat.de
Figure 4. Hydrogel actuation mechanism, origami, and applications. a) Hydrogel actuation mechanism: swelling, de-swelling. b) Reversible folding
and un-folding of a box with hydrogel at the hinges (Reproduced with permission.[148 ] Copyright 2011, American Chemical Society). c) Micro-scale
trilayer origami with complex deformation (Adapted with permission.[147 ] Copyright 2015, Wiley). d) Hydrogel gripper capable of locomotion and cell
capture/excision (Adapted with permission.[161 ] Copyright 2015, American Chemical Society). e) Hydrogel actuator with synergistic deformation and
color change (Adapted with permission.[152 ] Copyright 2018, Wiley). f) Hydrogel robot with octopus-like locomotion (Adapted with permission.[ 237 ]
Copyright 2021, Elsevier). g) Hydrogel metamaterials with negative swelling behavior (Adapted with permission.[159 ] Copyright 2018, the Authors, some
rights reserved; exclusive licensee AAAS).
materials with different stimuli-responsiveness. For example,
Zhang et al.[148 ] used a hinge consisting of a layer of pNI-
PAM hydrogel and a layer of low-density polyethylene (LDPE)
for different origami structures (Figure 4b). To simplify the
fabrication, hydrogel was only used in the patterned hinge re-
gions. They further loaded CNTs in the hydrogel to tune the re-
sponse time and to allow for near-IR irradiation-induced fold-
ing. Many other materials such as graphene oxide or reduced
graphene oxide,[152,153,228–230 ] MoS2,[ 154 ] and clay[151,157,228 ] have
also been incorporated with hydrogels to enhance the response
time and/or mechanical properties. It should be noted that other
various demonstrations of bilayer hydrogel origami have been
conducted by Ionov and coworkers, who have studied self-rolling
bilayer films[227 ] which can fold to 3D structures[149,216 ] such as
pyramids.
While the abovementioned bilayer origami is relatively simple,
multi-layered designs have also been used to enable more com-
plex hydrogel origami.[59,147 ] Na et al.[147 ] demonstrated micro-
scale origami with controlled mountain and valley folds with a
three-layer design: a thermoresponsive hydrogel layer between
two rigid polymer layers patterned with crease lines for either
mountain folds or valley folds. Upon immersion in water be-
low the LCST, the origami self-folded to its complex micro-scale
shape (Figure 4c), while it nearly reversibly returned to its origi-
nal configuration when heated above the LCST. They also demon-
strated the folding of a tessellation with 198 patterned creases
(Figure 4c). Although the origami shown is very impressive, it is
important to note that such complex self-folding origami can suf-
fer from mis-folding,[116,231 ] resulting in final configurations that
are different from the desired ones. To address this issue, the
same group used hydrogels with two different LCSTs, patterned
in a way such that the vertices of a complex origami design could
be pre-biased[232 ] toward certain directions, allowing for more re-
liable sequential folding of the origami creases to the desired final
3D shape.
As mentioned previously, hydrogels have been widely used for
biomedical applications. Self-rolled tubes for vascularization[233]
as well as numerous grippers[30,161,162,234 ] for drug delivery or
cell removal (biopsy) in surgical applications have been demon-
strated with hydrogels. Notably, Gracias and coworkers[161,162 ]
have created origami microgrippers by incorporating stiff poly-
mer panels with a thermoresponsive hydrogel substrate, where
the stiffness of the panels granted additional functionality to the
gripper, such as the possibility of cell excision. Grippers were
able to locomote[161 ] via the incorporation of magnetic particles
into the hydrogel layer, allowing remote control by a magnetic
Adv. Mater. 2023, 2302066 © 2023 Wiley-VCH GmbH
2302066 (8 of 35)
15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202302066 by Stanford University, Wiley Online Library on [12/12/2023]. 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 www.advmat.de
probe. Cell excision was accomplished by placing the grippers in
a warm environment with cell tissue. Starting in a closed con-
figuration, upon heating, the gripper opened up and folded in
the opposite direction, grasping a portion of tissue (Figure 4d)
and excising the cells from the tissue upon full closure of the
gripper. Additionally, drug delivery[162] from the grippers was
demonstrated by the incorporation of drugs into the layers of the
gripper. Other origami-like approaches to drug delivery[160] and
cell/particle encapsulation[163,164 ] using hydrogels have been ex-
plored as well.
Aside from biomedical applications, hydrogels are also perti-
nent to biomimetic applications, with some hydrogels exhibit-
ing synergistic fluorescent color and shape change.[152,235 ] Ma
et al.[152 ] designed a bilayer gripper (Figure 4e) with the bottom
side as a thermoresponsive hydrogel. At high temperatures, the
bottom layer expelled water and this led to downward bending,
exhibiting a fluorescent top layer under the green light. Upon a
decrease in temperature, however, the actuator would bend up-
ward and the color would change as a result. Another exciting
and recent example of biomimetic origami was the utilization
of hydrogel on a polydimethylsiloxane (PDMS) substrate with
microfluidic channels.[220 ] The device could respond to tempera-
ture, light, and humidity for an environmentally responsive actu-
ator with plant-like movements and photosynthesis capabilities.
Lastly, light-actuated hydrogel origami[230,236,237] has enabled soft
robots capable of crawler and octopus-like[237 ] motions. By utiliz-
ing bilayers composed of spiropyran functionalized with photo-
contracting and photo-expanding hydrogel, Li et al.[237 ] achieved
large bending deformation. By integrating the bilayers with ratch-
eted feet for anisotropic friction with a substrate, they achieved
biomimetic, octopus-like motions (Figure 4f) upon repeated re-
laxing and contraction of the bilayers.
Additional hydrogel origami applications entail metama-
terials with negative swelling enabled by positive-swelling
hydrogel.[159,238 ] Zhang et al.[159 ] utilized hydrogel to achieve
metamaterials with negative swelling (Figure 4g) by adjusting the
initial geometry of the metamaterial units. In changing the unit
geometries, metamaterials with expansion in one direction and
contraction in the orthogonal direction were achieved, demon-
strating interesting mechanical properties. Also, hydrogel for the
assembly of 3D, origami-like structures used for acoustic meta-
materials has been demonstrated by Deng et al.[158 ]
2.3. Liquid Crystal Elastomers (LCEs)
LCEs are a class of active materials that exhibit large,
reversible[239 ] deformation upon exposure to stimuli such
as heat,[240,241 ] light,[242,243 ] solvents,[ 244] electric,[ 245,246 ] or
magnetic[247 ] fields. LCEs are elastomers which have elon-
gated, anisotropic mesogens (typically composed of rigid
aromatic or cyclohexyl rings[248,249 ]) fixed to the polymer main
chain or sidechains by covalent bonds and crosslinks.[248 ] The
flexible nature of the elastomer allows for changes in the or-
dering of the mesogens upon stimuli, along with reversible
actuation.[250,251 ] Mesogens in LCEs are programmed to have
long-range orientational order described by a director, which
means that it has a high level of molecular order similar to that of
a crystalline solid, while maintaining fluidity similar to a viscous
liquid.[252 ] LCEs can exhibit mechanical, optical, electrical, and
magnetic anisotropic properties[248,253 ] in their ordered, or most
commonly “nematic”, state and isotropic behaviors in their dis-
ordered, or “isotropic”, state.[248 ] Due to their ease of reversible
deformation and large actuation strain, LCEs are used for a
number of applications, including artificial muscles[239,241,254 ]
and smart textiles.[255 ] Additionally, because the actuation of
LCEs is a phase transition process not governed by diffusion,
LCEs can actuate in several seconds, or even in a fraction of a
second.[181,256 ] The stiffness of LCE and the actuation stress, on
the other hand, usually range from hundreds of kPa to several
MPa.[241,256,257]
2.3.1. LCE Mechanism
The actuation of LCEs is due to the transition from a liquid
crystalline phase, such as nematic or smectic, to an isotropic
phase in response to a stimulus (typically temperature) that dis-
rupts mesogen alignment (Figure 5a). For most LCEs, there
exists a phase transition temperature referred to as the ne-
matic to isotropic phase transition temperature (TNI). Applica-
tion of an appropriate stimulus leads to a contraction paral-
lel to the director of the LCE.[239,258 ] Upon heating, polymer
chains gain entropy and become more coiled, causing mesogens
to lose their alignment and become randomly oriented, result-
ing in a macroscopic shrinkage of the LCE.[249,250,259 ] Thus, the
alignment and orientation of mesogens in LCEs are of critical
consideration in LCE applications. Previously, mesogens were
commonly aligned via surface rubbing[260 ] or electric fields,[261]
which are relatively tedious methods. More recently, a two-step
reaction method developed by Yakacki et al.[262 ] is often used,
which first involves a thiol-Michael addition “click” reaction
to form the polydomain LCE (or LCE without long-range or-
der), followed by alignment of mesogens by stretching, and fi-
nal locking of the alignment via photopolymerization reaction.
Planar nematic, twisted nematic, and splay are common liq-
uid crystal orientations for LCEs. Twisted and splay orientations
(Figure 5b)[263] involve variations in the director through the
thickness of the material, allowing for more complex deforma-
tion. In twisted and splay orientations, there exists a 90-degree
rotation of the director through the material thickness. Twisted
orientation involves an in-plane rotation of the director while
splayed is out-of-plane. When an appropriate stimulus is ap-
plied, mismatches in the expansion of the surfaces leads to bend-
ing/folding behaviors.[244,264 ] Because of this, different mesogen
orientations can be exploited for origami-like deformations of
materials.
2.3.2. LCE Origami
Thermo- and light- responsive LCEs comprise the majority of
LCE origami works, although it should be noted that humid-
ity and water-stimulated folding of LCE-based structures is also
achievable, often via bilayers,[265 ] alkaline,[174 ] or acidic[ 175] sur-
face treatment to enable asymmetric humidity response. White,
Ware, and collaborators have done a series of works related to
the folding of LCE. In a seminal work, they enabled the program-
ming of complex director orientations into thin LCE via surface
Adv. Mater. 2023, 2302066 © 2023 Wiley-VCH GmbH
2302066 (9 of 35)
15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202302066 by Stanford University, Wiley Online Library on [12/12/2023]. 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 www.advmat.de
Figure 5. LCE actuation mechanism and origami. a) Nematic (ordered) to isotropic (disordered) transition of LCEs. b) Macroscopic deformation of
LCE according to common mesogen alignments (Adapted with permission.[263 ] Copyright 2020, AIP Publishing). c) Optimal hinge design for LCE box
and folding of LCE upon heating (Adapted with permission.[171 ] Copyright 2015, The Royal Society of Chemistry). d) Deployment of reconfigurable
structure by actuation of LCE fibers (Reproduced with permission.[179 ] Copyright 2022, Wiley). e) Multimaterial origami airplane structure with LCE
actuators triggered by Joule heating (Adapted with permission.[033 ] Copyright 2017, The Royal Society of Chemistry). f) Pop-up of a liquid metal (LM)-
LCE composite triggered by induction heating (Adapted with permission from the authors).[181 ] g) LCE origami with synergistic shape and color change
(Reproduced with permission.[168 ] Copyright 2021, Wiley). h) Self-propelled LCE rolling robot (Adapted with permission.[172 ] Copyright 2019, AAAS).
alignment, in which local twisted nematic alignments were ef-
fectively defined as the origami hinges.[173 ] Upon heating, the
flat sheet would fold into Miura–ori. However, a downside was
the lack of control over folding upon heating. They later used
the same surface alignment technique and developed a topol-
ogy optimization[171 ] method to generate more reliable hinge
designs, with an aim to avoid anti-clastic, saddle-like curvature
seen in the actuation of twisted nematic LCEs.[264,266,267 ] Their
optimal hinge involved programmed triangular regions near the
substrate’s corners (Figure 5c-i) with approximately twisted ne-
matic order patterned through the thickness for nearly com-
plete closure of the box upon heating (Figure 5c-ii). They fur-
ther used LCE-CNT composites[176 ] as the hinges to enable a
light-activated LCE folding box that took advantage of CNT pho-
tothermal effects. Lastly, Donovan et al.[ 268] demonstrated a long-
lasting folded LCE box by incorporating o-fluorinated azoben-
zene into the polymer network. The trans-to-cis isomerization
of the o-fluorinated chromophores resulted in directional strains
on the LCE and allowed for the macroscopic deformation, which
could last for multiple days. This type of prolonged deformation
of LCE can be beneficial, as a constant stimulus would typically
be needed for LCE to maintain its deformation, which may un-
necessarily consume energy. This topic will be discussed further
in Section 2.5.
Other origami-like folding of flat sheets to 3D shapes has been
achieved by introducing complex LCE director distributions re-
alized by photoaligned substrates or molds with micrometer-
scale channels.[269–272 ] Another more simplified approach was to
Adv. Mater. 2023, 2302066 © 2023 Wiley-VCH GmbH
2302066 (10 of 35)
15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202302066 by Stanford University, Wiley Online Library on [12/12/2023]. 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 www.advmat.de
program the folding of bilayer origami LCE sheets by spraying
LCE onto thermoplastic sheets, as demonstrated by Verpaalen
et al.[177 ] Upon thermally shaping and fixing the sheets, sharp,
origami-like folds could be introduced. These thin bilayer strips
were actuated by UV and blue light and simple deformation from
accordion to flower-like shape was demonstrated.
Instead of designing LCE origami based on entire sheets of
LCE, axially aligned LCE fibers, or strips of LCE that contract
along their length when actuated, can also be selectively placed
among origami structures to generate folding.[178–180,33 ] As seen
in Figure 5d, Peng et al.[179] printed LCE fibers onto a square-
shaped reconfigurable structure. When the structure was loaded,
it fell to a parallelogram-like shape that could be deployed back
to the initial state upon heating the fibers. In their work, they
also demonstrated deployment of a printed tensegrity structure
and reconfiguration of an LCE lattice pyramid. An additional ad-
vantage of the use of LCE fibers/strips compared to LCE sheets
is that they can be selectively actuated by methods such as Joule
heating, which can enable more complicated folding routines. By
using LCE fibers as the actuators, Yuan et al.[33] demonstrated
a variety of multimaterial origami structures composed of an
elastomeric material (Tangoblack by Stratasys (Rehovot, Israel)),
a rigid polymer (Verowhite by Stratasys), and a conductive ink.
They obtained an inactive substrate by inkjet printing and then
printed conductive ink onto this substrate via direct ink writing
(DIW), after which they incorporated the LCE strips into the pre-
determined positions atop the ink. Under an applied current, the
conductive ink was Joule heated, leading to the actuation of the
origami. Once the current was ceased, the LCE cooled down and
the actuator gradually returned to its original flat state. An air-
plane actuator, as seen in Figure 5e, utilized LCE fibers on the top
and bottom sides to serve as actuators for upward and downward
bending, respectively. Miura–ori, as well as a sequential folding
box were also demonstrated with this method.
While sequential or selective actuation (and thus folding) of
LCE origami can be obtained from Joule heating, a limitation of
this is that wires are often needed to supply current, which com-
plicates the structure. In a recent work, Maurin et al.[181 ] devel-
oped liquid metal (LM)-LCE composites that could be remotely
heated by induction under alternating magnetic fields. The heat-
ing of LCE was concentrated in regions where the LM was lo-
cated, and the LM could be patterned arbitrarily due to the LM-
spraying method that was used. By varying the quantity of LM
and the LM pattern, as well as the alignment of LCE and the mag-
nitude of the applied magnetic field, the actuated shape of the
LCE could be programmed. As shown in Figure 5f, an LCE sam-
ple with circumferential alignment (which buckles out of plane
when actuated[173 ]) and LM throughout the composite sequen-
tially “pops up” under an increasing magnetic field, where the
centermost LCE buckles first, followed by the outermost LCE. In
their work, the LM-LCE composite was also used to design sea-
turtle-inspired actuator “fins” for a biomimetic turtle robot.
Efforts have also been taken to realize color change during
the actuation of LCE origami.[165,168 ] Huang et al.[ 168] recently
presented luminescent LCE origami by incorporating fluores-
cent and color-changing agents into the LCE matrix, includ-
ing a blue-green fluorescent moiety tetraphenylethene, a pho-
tochromic moiety spiropyran, and an NIR photothermal dye
YHD796. NIR light was used as a heat source to facilitate the
deformation of the LCE, while UV and visible light were used
for reversible color change between green-blue and red at folded
states and dull red at deployed states (Figure 5g). Both Miura–
ori and Yoshimura origami were demonstrated by introducing
folds mechanically, creating a gradient of mechanical stress that
allowed for reversible folding and unfolding. By controlling the
applied light, synergistic shape and color change of the origami
was achieved, enabling biomimetic behaviors such as camouflag-
ing.
LCE-based active origami has also been used to generate lo-
comotion for robotics applications. Kotikian et al.[172 ] assembled
multiple different examples of LCE origami, taking advantage
of thermoresponsive LCEs with different nematic to isotropic
transition temperatures. They utilized rigid acrylate panels with
LCE bilayer hinges. The hinges involved two LCE layers printed
orthogonal to one another to enable large bending curvatu res
(up to 180 degrees). To achieve sequential folding, they strate-
gically used hinges with LCEs of either high TNI or low TNI
throughout their designs, and demonstrated sequentially fold-
able structures, as well as a self-propelling, rolling robot, as
shown in Figure 5h. This robot utilizes low TNI (blue) hinges
to gradually fold the body into a pentagonal prism and high
TNI (orange) hinges to propel the robot forward, allowing for
continual rolling motion. Upon activation by a hot plate, the
hinges provide torques which sustain the forward motion of
the robot, even after undergoing a full revolution. Other works
have demonstrated locomotion via LCE actuation for crawling[273]
origami robots. Additional exciting LCE origami applications
include thermoresponsive LCE kirigami metasurfaces[274 ] and
LCE metamaterials demonstrated as shrinkable patches for skin
regeneration.[275 ]
2.4. Magnetic Soft Materials (MSMs)
MSMs are a class of active materials that permit fast, reversible,
and untethered deformation under an applied external mag-
netic field. MSMs for active origami are obtainable by various
strategies, including embedding magnetic particles into a poly-
meric matrix or coating structures with magnetically responsive
metal particles.[182 ] The magnetic fillers or coatings are the essen-
tial components of MSMs for generating mechanical load (body
torque) or heat under the external magnetic field, realizing the
shape morphing of active origami. Various types of magnetic
fillers, with ferromagnetic (soft-magnetic, hard-magnetic), or su-
perparamagnetic properties, behave differently under a magnetic
field and require specific control strategies. Magnetic fillers can
be integrated with inactive soft polymers or functional active ma-
trices of SMP, LCE, hydrogel, dynamic polymer, etc. Due to the
wide range of materials that MSMs can be based on, the stiff-
ness and actuation stress of these materials varies accordingly.
In addition, MSMs can actuate rapidly, from tens of microsec-
onds to seconds.[37,38,276 ] This section focuses on how magnetic
fillers and magnetic control facilitate reconfiguration and loco-
motion of active origami systems. Active origami that incorpo-
rates magnetic fillers with specialized polymeric matrices, for
example, inductive heating of SMP by embedding superpara-
magnetic particles, is discussed in the corresponding material
section.
Adv. Mater. 2023, 2302066 © 2023 Wiley-VCH GmbH
2302066 (11 of 35)
15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202302066 by Stanford University, Wiley Online Library on [12/12/2023]. 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 www.advmat.de
2.4.1. Magnetic Actuation Mechanism
The certain magnetic fillers and coatings used in MSMs de-
cide their actuation mechanisms. Soft-magnetic materials (iron,
nickel, iron–nickel alloys) cannot retain strong remanent magne-
tization and can be easily de-magnetized and re-magnetized by
an applied magnetic field. Because of this, these materials rely
on a gradient magnetic field to generate force. Therefore, struc-
tures and materials embedded/coated with soft-magnetic mate-
rials require a gradient magnetic field in order to be actuated. A
gradient magnetic field can simply be obtained by a permanent
magnet.[277 ] A moving magnet can be used for locomotion and
navigation of MSM active origami systems due to the attraction
force toward the magnet. However, the magnetic field gradient is
typically nonuniform around a magnet, making the programma-
bility and accurate control of soft-MSMs relatively limited. On
the other hand, hard-magnetic materials such as chromium diox-
ide (CrO2), neodymium–iron–boron (NdFeB), and hard ferrite
can retain strong remanent magnetization after being magne-
tized under a strong magnetic field (1.5 to 2 T). When a uni-
form magnetic field is applied, a torque can be generated to ro-
tate the magnetic materials so that their magnetization is aligned
with the applied magnetic field. Since a uniform magnetic field
can be generated by a pair of Helmholtz coils, the use of hard-
magnetic particles as fillers is thus promising, as more accurate
programmability and control can be obtained. Therefore, this sec-
tion mainly discusses MSMs using hard-magnetic materials (re-
ferred to as hard-MSMs). It should be noted that hard-MSMs can
be re-magnetized by either heating them to above their Curie
temperature followed by applying a strong magnetic field (1.5–2
T) or by directly applying a strong magnetic field. As illustrated by
a beam with the programmed magnetization along its axial direc-
tion (Figure 6a), micro-torques are exerted on the hard-magnetic
particles when their magnetization is not aligned with the applied
magnetic field B. These micro-torques are transmitted to the soft
matrix and lead to a bending deformation under an upward mag-
netic field.[278 ] Another prominent feature of hard-MSMs is the
flexibility to pattern magnetization distributions on demand, per-
mitting complex shape reconfiguration of active origami. Kim
et al.[37] developed a DIW printing method of hard-MSMs, which
programs the magnetization along the filament printing direc-
tion (more details discussed in Section 3.2.1). This enabled the
fabrication of magnetically responsive origami, for example, the
Miura–ori pattern in Figure 1c as well as the hexapedal structure
with various magnetizations shown in Figure 6b, which realized
wirelessly controlled, fast and reversible contraction under an ap-
plied magnetic field.
2.4.2. MSM Origami
Magnetic actuation has been demonstrated for effective fold-
ing of origami structures on a variety of scales, based on dif-
ferent structural fabrication and magnetization strategies. The
patterned magnetization, the external magnetic field, and the
structural design (panels and hinges) determine the origami fold-
ing. Cui et al.[184 ] encoded magnetization of microscale origami
structures by fabricating nanomagnet arrays on submicron pan-
els with electron beam lithography. As illustrated by the crane
microrobot in Figure 6c, the panels were assigned with dis-
tributed magnetizations and the flat sheet reconfigured to a “flap-
ping configuration” under the applied magnetic field. Note that
the designed nanomagnet array possessed anisotropic proper-
ties, enabling reprogrammable magnetization by a specially de-
signed sequence of magnetic fields for a “hovering” mode. The
re-programmability of magnetic materials allows for adaptable
shape changing and tunable properties even after material fabri-
cation and thus enhances the functionality of active origami for
applications such as soft robotics, metamaterials, and reconfig-
urable structures. Alapan et al.[189 ] developed a reprogrammable
metamaterial (Figure 6d) using a soft composite embedded with
CrO2particles, which have a relatively low Curie temperature of
118°C. Specific panels were selectively heated >150°C (using a
laser) while a moderate 15 mT encoding magnetic field was ap-
plied, which reprogrammed the magnetization pattern and main-
tained 90% of the initial magnetization magnitude. Alternatively,
Song et al.[279 ] demonstrated reprogrammable origami sheets
by first integrating NdFeB particles and phase change material
(oligomeric polyethylene glycol) into microspheres, and then em-
bedding the magnetic microspheres into the elastomeric matrix
as shown in Figure 6e. Upon heating, the phase change material
in the microspheres melted and the magnetized magnetic parti-
cles rotated based on the applied magnetic field direction. After
cooling down, the phase change material solidified, locking the
alignment of magnetic particles and the corresponding magneti-
zation. Through this strategy, an origami sheet was demonstrated
with various transformations into desired shapes under magnetic
actuation (Figure 6e). Other reprogrammable MSMs with further
integrated functions are also available based on covalent adapt-
able networks due to the nature of the polymeric matrix, which
will be discussed in the next section.
Apart from shape morphing structures and metamaterials,
MSM origami systems are also widely used for soft robotics.
Gracias and collaborators[186] have done extensive work on un-
tethered origami microgrippers (Figure 6f). The microgrippers
with magnetic coating allowed for wirelessly controlled navi-
gation in complex environments such as the stomach and in-
testines. In addition, the reversible opening and closing abili-
ties of the microgrippers under different stimuli including pH
change,[30] temperature,[161,280] chemicals,[186 ] etc., have been
demonstrated for functions of object grasping, cargo transporta-
tion, and biopsy. Inspired by jellyfish, Ren et al.[183 ] demon-
strated a hard-MSM swimming robot with an engineered hinge
as shown in Figure 6g. By manipulating the applied magnetic
field profiles, the swimming robot had different swimming mo-
tion modes for multiple functionalities such as the mixing of
chemicals and transporting of objects. Recently, Wu et al.[185]
demonstrated a magnetic origami robotic arm that integrated
multimodal deformations of stretching, folding, omnidirectional
bending, and multi-axis twisting into one system as illustrated
in Figure 6h. By integrating the distributed magnetizations and
precise magnetic control into serially arranged Kresling origami,
the robotic arm illustrated sophisticated motions that mimic an
octopus arm. In addition, a small-scale origami crawler was also
realized by Ze at al.[75] via simultaneous contraction of four Kres-
ling units (Figure 6i). Here, the magnetic actuation separated
the power source and control system out of the robot, enabling
the miniaturization of the origami structure. The crawler was
Adv. Mater. 2023, 2302066 © 2023 Wiley-VCH GmbH
2302066 (12 of 35)
15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202302066 by Stanford University, Wiley Online Library on [12/12/2023]. 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 www.advmat.de
Figure 6. MSM actuation mechanisms, origami, and applications. a) Actuation mechanism of hard-MSMs. b) Patterned magnetization of hard-
MSMs for complex shape morphing (Adapted with permission.[37] Copyright 2018, Springer Nature). c) Microrobot with reprogrammable shape mor-
phing (Adapted with permission.[184 ] Copyright 2019, Springer Nature). d) Reprogrammable metamaterial through coupled magnetic and thermal
loadings (Adapted with permission.[189 ] Copyright 2020, the Authors, some rights reserved; exclusive licensee AAAS). e) Reprogrammable origami
sheets by integrating phase change materials with magnetic particles in a soft matrix (Adapted with permission.[279] Copyright 2020, American Chem-
ical Society). f) Untethered microgripper capable of navigation in complex environments and grasping of objects (Adapted with permission.[186]
Copyright 2009, National Academy of Sciences). g) Soft jellyfish robot with multiple functionalities such as chemical mixing and object transporta-
tion (Adapted with permission.[183 ] Copyright 2019, Springer Nature). h) Robotic arm with integrated multimodal deformations of stretching, fold-
ing, omnidirectional bending, and multi-axis twisting (Adapted with permission.[185 ] Copyright 2021, National Academy of Sciences). i) Soft robotic
origami crawler with drug delivery capability (Adapted with permission.[75] Copyright 2022, the Authors, some rights reserved; exclusive licensee
AAAS).
Adv. Mater. 2023, 2302066 © 2023 Wiley-VCH GmbH
2302066 (13 of 35)
15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202302066 by Stanford University, Wiley Online Library on [12/12/2023]. 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 www.advmat.de
Figure 7. CAN mechanism and origami. a) Bond exchange reaction schematic and typical CAN properties of self-healing, plasticity, andreprogramming.
b) Elasticity and plasticity of an SMP with dynamic covalent bonds and reprogrammable permanent shapes. Adapted with permission.[292 ] Copyright
2016, AAAS). c) Shape memory and welding of materials for sequentially foldable structures (Adapted with permission.[298 ] Copyright 2016, Wiley). d)
LCE with multi-shape memory capabilities controlled by light-activated bond exchange and heat for shape deformation (Adapted with permission.[169 ]
Copyright 2018, the Authors, some rights reserved; exclusive licensee AAAS). e) LCE with photo-patterned, de-crosslinked regions as hinges (Adapted with
permission.[170 ] Copyright 2019, Wiley). f) Magnetic dynamic polymers with reprogrammable magnetization (Adapted with permission.[307] Copyright
2021, Wiley).
explored for multiple functionalities such as movement in a con-
fined space, drug storage, and drug release. An additional notable
work based on Kresling origami was a soft amphibious robot with
swimming, jumping, and cargo transportation functionalities.[31]
2.5. Covalent Adaptable Network (CAN) Polymers
Covalent adaptable network (CAN) polymers[281 ] are crosslinked
polymers with dynamic covalent bonds that can reversibly break
and re-form while maintaining network integrity (Figure 7a).
They are also often referred to as dynamic covalent polymer net-
works (DCPNs),[282 ] or vitrimers.[283 ] Under an applied stimu-
lus, often heat,[283–285 ] light,[286 ] or solvent,[287 ] the so-called bond
exchange reactions (BERs) become active in the network (see
Figure 7a, left), which lead to interesting merits such as plastic-
ity, recyclability, self-healing, and ease of reprogramming, equip-
ping CANs with specialized functionalities as compared to the
aforementioned materials (Figure 7a, right).[281,282] These have
led to applications such as healable sensors[288 ] and healable tri-
boelectric nanogenerators.[289 ] As we will discuss, dynamic cova-
lent chemistry (DCC) can be incorporated with active materials
to achieve active origami with outstanding properties.
2.5.1. CAN Mechanism
Typical CAN polymers involve BERs that enable the rearrange-
ment of networks. Common chemistries that possess BERs in-
clude transesterification, disulfide exchange, imine exchange,
Diels–Alder (D–A), etc. Generally, the reactions that allow for
adaptability of bonds are dissociative or associative.[290 ] Dis-
sociative (such as D–A) networks involve covalent bonds that
are not necessarily re-formed upon breaking, allowing for sig-
nificant decrease in the number of crosslinks in the net-
work. For associative (such as the abovementioned exchange
reactions except for D–A) networks, the overall number of
crosslinks in the network remains the same.[281 ] Upon acti-
vation of BERs, topological rearrangement of the network oc-
curs, where nearby bonds break and re-form, which can allow
Adv. Mater. 2023, 2302066 © 2023 Wiley-VCH GmbH
2302066 (14 of 35)
15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202302066 by Stanford University, Wiley Online Library on [12/12/2023]. 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 www.advmat.de
for plastic deformation of the material[282 ] and shape repro-
gramming. In addition, when BERs cross the interface be-
tween two CAN polymers, welding, reprocessing and recycling
are enabled. Figure 7b shows how BERs can be used together
with shape memory effects to achieve diverse programmability
and shape change. The shape memory effect allows the poly-
mer sheet to be programmed into a temporary shape and to
be elastically recovered to the permanent shape (Figure 7b-i)
while the permanent shape can be plastically reprogrammed by
BERs (Figure 7b-ii). Since the chemical reaction kinetics typically
follow Arrhenius law,[291] thermoresponsive BERs only become
fast enough (or active) to allow appreciable changes at high tem-
peratures (TBER). In addition, since BERs require macromolec-
ular chains to have enough mobility for exchange reactions, the
temperature for BERs to be active is typically higher than Tg.This
permits the possibility to combine the merits of BERs with shape
memory effects.
2.5.2. CAN Origami
Xie and coworkers have pioneered multiple notable works on
CANs paired with SMPs to create active origami. In their ini-
tial works,[292,293 ] they achieved elasticity and plasticity within the
same polymer network. Plasticity could be induced by transes-
terification BERs, allowing for reprogramming of complex 3D
permanent shapes upon heating to TBER. By deforming the sheet
while above its Ttrans (but below TBER) followed by cooling, tem-
porary origami shapes could be imparted (Figure 7b-iii). Upon
heating, the original origami shape would be recovered. By rais-
ing the temperature close to or above TBER, the permanent shape
could be reprogrammed. In a later work, they again used plas-
ticity and elasticity concepts, but to demonstrate the ability of
2D origami substrates to transform to 3D origami[294 ] via plastic-
ity, while reversible elastic deformation could occur at this state,
showing value for deployable structures that require subsequent
localized actuation. Plasticity of SMP CAN origami has also been
explored by other groups recently.[295,296]
Pei et al. incorporated reversible links into LCEs to achieve
easy processing and alignment after the initial cross-linking, en-
abling reprogramming.[297 ] They later utilized this concept to
demonstrate CAN active origami.[298–302 ] Via transesterification
reaction, they showed that light-activated CNT-dispersed LCE
could be aligned spatio-selectively, in which local parts of the
material could be photothermally heated until transesterification
was activated, after which those parts of the LCE actuator could
be aligned via stretching.[299 ] They presented another approach
that used CANs for jigsaw-like assembly of multimaterial origami
substrates.[298 ] Three CANs with different TBERswereusedto
weld the materials via hot press to flat crosses of varying material
arrangement. Upon programming a box shape and subsequent
heating, un-folding of multimaterial boxes (Figure 7c) could oc-
cur in different sequential orders due to the order in which the
material phase transitions occurred. They additionally demon-
strated LCE CANs activated by solvent,[301,302 ] allowing for actua-
tion of LCE origami, such as Miura–ori, 3D kirigami structures,
and flower-like actuators at relatively moderate temperatures.
Bowman and coworkers have made significant advance-
ments in the understanding of covalent adaptable networks[281 ]
and developed the first known light-triggered bond exchange
reaction.[286 ] Unlike photothermal effects, they used light to di-
rectly trigger bond exchange reaction (or reversible addition-
fragmentation chain transfer (RAFT)).[286 ] The light-triggered
BER was later used to demonstrate photo-origami[34] and has
demonstrated active origami with exciting properties.[169,303 ] De-
picted in Figure 7d, they developed LCEs which have light-
activated BERs that allowed for separate control of the LCE de-
formation and alignment.[169 ] Here, the RAFT functionality was
incorporated into the polymer backbone and enabled bond ex-
change of allyl sulfides within the network. The light-induced
BER allowed for stabilized alignment of LCEs in a strained state,
resulting in stable, permanent origami shapes upon removal of
strain. When heated, these structures deformed to flat sheets
and returned to the 3D shape upon cooling. In this way, re-
versible folding of Miura–ori was achieved. Additionally, they
demonstrated transformation between three distinct shapes, as
shown in Figure 7d, in which the 3D crane was the permanent
shape programmed via BER, the flat crane was the isotropic
state, and a third temporary state was able to be temporarily
programmed via thermal quenching. Li et al. also demonstrated
triple shape memory of LCE boxes,[304,305 ] taking advantage of
both the glass and isotropic phase transition temperatures of the
LCE.
Active CAN LCE origami has also been demonstrated by Zhao
and coworkers,[170,306 ] who utilized anthracene moieties with liq-
uid crystals, achieving light-controlled cleavage (de-crosslinking)
and dimerization (crosslinking) under two UV lights of differ-
ent wavelengths.170 In this way, de-crosslinked, inactive domains
could be introduced through the thickness of an LCE sheet,
which effectively created a bilayer of the original LCE and the
newly de-crosslinked LCE. The de-crosslinked regions acted as
flexible hinges and caused curvature of the sheet at ambient tem-
peratures, with the extent of bending controllable by the extent of
decrosslinking that was programmed through the thickness. Re-
gions with thicker inactive domains resulted in smaller bending
angles toward the de-crosslinked side. Upon heating the sheet,
contraction of the crosslinked regions led to an overall flatten-
ing of the sheet and temporary loss of the origami shape. They
used the same sheet and progressively patterned more inactive
domains (See Figure 7e), to achieve origami shapes such as an
airplane, bull, and frog.
CANs were also recently demonstrated with MSMs. Kuang
et al.[307 ] introduced magnetic dynamic polymers by embedding
hard magnetic NdFeB particles in a D–A polymer matrix. At ele-
vated temperatures, cleavage of dynamic linkages was triggered,
allowing for reprogramming of the magnetization under rela-
tively low magnetic fields. They demonstrated this with a foldable
array of magnetic dynamic polymer squares, initially with hori-
zontal magnetizations opposite to those in neighboring columns
(see Figure 7f, left). This type of magnetization resulted in a
3D “W” shape upon an applied downward magnetic field. With
the use of a photomask and IR light for photothermal heating,
the magnetizations of squares were selectively reprogrammed to
four different diagonal magnetizations (Figure 7f, right). They
also demonstrated the use of simultaneous applied magnetic
field and bond exchange to obtain stress-free bistable and mul-
tistable kirigami structures that would otherwise be difficult to
fabricate without BER.
Adv. Mater. 2023, 2302066 © 2023 Wiley-VCH GmbH
2302066 (15 of 35)
15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202302066 by Stanford University, Wiley Online Library on [12/12/2023]. 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 www.advmat.de
Figure 8. Additional materials for active origami. a) Hydrophilic and hydrophobic graphene-based origami, which absorbs and desorbs water molecules
for actuation (Adapted with permission.[312 ] Copyright 2015, AAAS). b) PEDOT:PSS-based origami which uses origami multistability to store mechanical
information (Adapted with permission.[076 ] Copyright 2018, National Academy of Sciences). c) Origami with thermally expanding microspheres which
result in bending of structures upon heating (Adapted with permission.[322 ] Copyright 2018, Wiley). d) Actuation of SMA spring by heating and adaptive
shading device which reduces its surface coverage when springs are heated (Adapted with permission.[017 ] Copyright 2018, ASCE). e) Active origami
with dielectric fluid at hinges to actuate wings of a crane under an applied voltage (Adapted with permission.[336 ] Copyright 2018, AAAS).
Although not thoroughly covered in this section, DCC can im-
part shape memory capabilities on hydrogels as well. Chen et al.
have recently developed fluorescent hydrogel origami, which uti-
lizes dynamic covalent bonding to enable shape memory func-
tion of hydrogel and dual encryption applications. Messages pat-
terned on the hydrogel could be decrypted upon shape recovery
and UV light illumination.[308,309 ] Generally, CANs enable unique
properties of active origami that are otherwise difficult to achieve
in SMPs, hydrogels, LCEs, or MSMs without DCC.
2.6. Other Materials
While numerous active origami works have been done with
SMPs, hydrogels, LCEs, and MSMs, there are additional stimuli-
responsive materials used for active origami worth highlighting
here. For simplicity, the materials discussed in this section are
categorized into the general groups of hygroscopic, thermore-
sponsive, and electroactive materials.
Hygroscopic materials are those that can absorb or desorb
water vapor and will thus change shape in response to varia-
tions in humidity. These materials are similar to hydrogels in
mechanism, however, unlike hydrogels, they do not require an
aqueous environment to operate. In an early study, Okuzaki
et al.[310 ] found that, for accordion-like origami actuators with
polypyrrole films, sorption and desorption of vapor drastically
changed the elastic modulus of the polypyrrole material, result-
ing in reversible strains of nearly 150%. Other materials, such
as GO (graphene oxide), are hydrophilic and can adsorb water
molecules.[311 ] As seen in Figure 8a, Mu et al.[312 ] developed a
graphene-based “paper” that was composed of a hydrophobic
rGO (reduced graphene oxide) layer and a hydrophilic GO-
polydopamine (GO-PDA) layer. They fabricated a flat cross
substrate with patterned GO-PDA hinges. Upon the photoin-
duced heat of NIR light irradiation, the GO-PDA regions released
water by desorption, causing a contraction of the layer and lead-
ing to upward bending of the rGO outer panels, which was
quickly recovered back to a flat state by water adsorption upon
removal of the light. PDA along with PNIPAM have also been
functionalized as hinges on graphene origami substrates.[313 ] Cai
et al.[314 ] have demonstrated the humidity response of MXene
cellulose composites for actuation of bilayer origami structures
with polycarbonate membranes. Other membranes such as the
porous cationic poly(ionic liquid) demonstrated by Zhao et al.[315 ]
exhibited rapid folding and un-folding in response to humidity
changes in the presence of organic solvents such as acetone.
Another common material used for hygroscopic response
is poly(3,4-ethylenedioxythiophene) polystyrene sulfonate
Adv. Mater. 2023, 2302066 © 2023 Wiley-VCH GmbH
2302066 (16 of 35)
15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202302066 by Stanford University, Wiley Online Library on [12/12/2023]. 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 www.advmat.de
(PEDOT:PSS), a conductive polymer used for a myriad of elec-
tronic devices.[316 ] It has been used to create active origami
structures.[76,317 ] For example, Treml et al.[76] used PEDOT:PSS
as the active material for origami logic structures, in which
PEDOT:PSS was used with polypropylene to create humidity-
responsive waterbomb origami films with bistability or multista-
bility (Figure 8b).
While thermal expansion mismatch in bilayer or multilayer
systems is often used to drive the shape change of active
origami, heat has been used in other ways to actuate origami.
For example, as temperature increases, polymers tend to soften,
which has been taken advantage of for origami structures, such
as the heating of polymer at pre-stressed metallic hinges of
microgrippers,[186,318 ] polymer sheets with reduced-thickness pat-
terned hinges and stretched rubber bands which actuate struc-
tures upon hinge softening,[319 ] as well as the folding of graphene
polyhedrons as a result of melting its polymer hinges.[320 ] In-
stances of the incorporation of graphene with bilayers in ac-
tive origami have been demonstrated.[321,322 ] For example, Tang
et al.[322 ] embedded thermally expanding microspheres (TEM)
into rGO/PDMS bilayers (Figure 8c). The microspheres, which
were heated via photothermal effect by the rGO upon exposure
to light, were liquid hydrocarbons encapsulated by thermoplastic
shells. When heated, the shell softened and the hydrocarbon tran-
sitioned to a gas, enabling an expansion of the spheres which re-
mained after the outer shell hardened upon cooling. This expan-
sion of the rGO/TEM portion of the bilayer caused irreversible
bending toward the PDMS.
Shape memory alloys (SMAs) are an additional type of ther-
moresponsive materials that have been used often in active
origami applications.[323,324 ] As the name indicates, SMAs exhibit
a shape memory effect similar to that of SMPs (but note that
SMAs had been used and developed prior to SMPs). In short,
these alloys can undergo a reversible phase transformation be-
tween martensite and austenite phases during which the crys-
tal structures change.[325 ] Due to their ability to recover their
shape even when under appreciable loads, SMAs have relatively
high actuation energy density as compared to other stimuli-
responsive materials and can be advantageous in this way. Note
that the allowable strain of SMAs varies by alloy, but is gener-
ally limited to <8%.[325,326 ] Different SMA waterbomb-based de-
signs with reduced-thickness hinges have been used to demon-
strate biomedical stents capable of coupled radial and axial[12]
or circumferential[327 ] deployment. SMAs have also been used
for the deployment of waterbomb-based transformable wheels,
where SMA wire contracts radially when heated, allowing for
reconfiguration of the wheels.[328 ] Additional works have used
SMA wires/springs as actuators for sunlight-adaptive Ron Resch
(Figure 8d)[17] or waterbomb[329 ] origami-based shading devices.
Additionally, SMA has been used toward robotics applications,
for instance the use of SMA wires for actuation of a peristaltic
origami crawler robot.[330 ] It should be noted that other impres-
sive examples of metal-based origami[331–333 ] include the surface
oxidation of platinum films[334 ] and metallic bilayers.[335 ]
Dielectric materials are also used to create active origami,
or so-called electroactive origami. Common dielectric materials
involve fluids, for example, dielectric liquid can be displaced
upon an applied voltage and subsequently drive the movement of
panels.[336,337 ] Figure 8e shows an example by Taghavi et al.,[336 ]
in which dielectric liquid was applied at the hinge of two op-
positely charged panels. Upon an applied voltage, the electro-
static attraction drew the panels closer to one another, resulting
in folding. With this concept, they demonstrated an accordion-
like artificial muscle and a flapping origami crane. Dielectric elas-
tomers can reduce in thickness and increase in area when a volt-
age is applied to electrodes across the thickness direction[338 ] and
can be used for simple motions such as bending,[339–342 ] as well
as for more complex origami.[272,343–345 ] Recently, Sun et al.[346]
utilized origami systems with elastomer between carbon grease
electrodes. The initial 3D structures (via pre-stretching) were ac-
tuated by applying a voltage (enabling electrostatic force between
electrodes), which caused the dielectric elastomer to expand, un-
folding the structure. The applied voltage could selectively actuate
the hinges individually, enabling boxes and pyramids as well as
gripper functionality.
Lastly, ceramics are also suitable materials for active origami.
Several works have utilized inactive ceramics, such as elastomer-
derived ceramics[347,348 ] for origami demonstrations, while others
use active materials such as piezoelectric ceramics. Under an ap-
plied electric field, piezoelectric ceramic materials such as PZT
(lead zirconate titanate) can expand in one direction.[349 ] By plac-
ing PZT between electrodes, voltage-driven actuation of the ce-
ramic can be achieved, which has been used for robotics[350–352 ]
and active origami applications.[23,353 ] Specifically, Suzuki and
Wood[23] have demonstrated the use of PZT as linear actuators
for an origami-inspired surgical robot.
3. Fabrication of Active Origami
As active origami involves the shape evolution of materials, the
fabrication and patterning of materials that facilitate complex de-
formations are of great importance. Here, we discuss the most
common methods used to fabricate active origami, including
molding and 3D/4D printing, as well as techniques used for mi-
cro and nano scale active origami. Table 2 provides a brief sum-
mary of the common fabrication methods.
3.1. Molding
Molding is a common method for fabrication of soft materials
in which precursor materials are poured into a mold and cured,
resulting in structures that take on the inner cavity of the mold.
Molding has been used in fabrication of SMPs, hydrogels, LCEs,
MSMs, and CANs. For SMPs, molding can initially be used to
obtain the “permanent” shape of the material, while additional
molds or methods for shape fixation can be used to program tem-
porary shapes. Fabrication of hydrogel structures also commonly
utilizes molding,[148,154,354,355 ] where multi-step processes can be
used to create bilayers or in which photomasks are used to in-
troduce crosslinking densities. Molding of LCEs[168,169 ] has been
used to introduce origami patterns into LCE sheets. Intricate
molds with micro-scale features such as microchannels[269,271 ]
have been used, which leads to alignment of the mesogens along
the channels to produce films with complex director fields. For
cured MSMs with hard-magnetic particles such as NdFeB, the
material can be placed in fixtures which configure the material
Adv. Mater. 2023, 2302066 © 2023 Wiley-VCH GmbH
2302066 (17 of 35)
15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202302066 by Stanford University, Wiley Online Library on [12/12/2023]. 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 www.advmat.de
Table 2. Summary of fabrication methods.
Method Material properties Active materials used Pros Cons
Molding Photo or thermally curable SMPs, Hydrogels,
LCEs, MSMs
Inexpensive Low structural complexity;
requires mold for each
new design
Direct ink writing (DIW) Viscous ink, shear-thinning SMPs,
Hydrogels, LCEs,
MSMs
Inexpensive; more material
choices
Low resolution;
slow
Fused filament fabrication
(FFF)
Thermoplastic SMPs,
Hydrogels, MSMs
Inexpensive Low resolution;
slow
Digital light processing (DLP) Photocurable, liquid resin SMPs,
Hydrogels, LCEs,
MSMs
Fast (cures entire layer) Multimaterial printing
challenging
Two-photon polymerization
(TPP)
Photocurable via two-photon
absorption
Hydrogels, LCEs,
MSMs
High resolution Difficult to scale up
Inkjet Photocurable liquid ink SMPs, Hydrogels Fast; high resolution;
multimaterial capability
Expensive; limited material
choice
to a desired actuation shape. The material can then be magne-
tized by applying a large magnetic field, imparting a magnetiza-
tion distribution in the MSM[73,356,357 ] so that it would later config-
ure to that shape under an appropriate magnetic field. It should
be noted that CANs can be fabricated in a similar way to that of
their matrix material, but with special considerations to the ma-
terial’s chemistry.
3.2. 3D/4D Printing
3D printing, or additive manufacturing (AM), has enabled the
rapid manufacture of custom designed parts. These techniques
have since been expanded to stimuli-responsive/shape-changing
materials to bring forth 4D printing,[358,359 ] or the printing of
structures whose shape, properties, or functions can evolve
over time. 4D printing has been widely adopted for the fab-
rication of complex, shape-evolving structures and materials.
Here, we will briefly discuss the 3D/4D printing methods com-
monly used to fabricate active origami, including extrusion-based
printing such as direct ink writing and fused filament fabrica-
tion, the resin-vat based digital light processing, inkjet print-
ing, as well as two-photon polymerization for high-resolution
printing.
3.2.1. Extrusion-Based Printing
Extrusion-based printing includes direct ink writing (DIW) and
fused filament fabrication (FFF; often called fused deposition
modeling (FDM)). These two methods share many common fea-
tures, such as they both use nozzles to write lines to form layers
and eventually form 3D solids. Both methods have been popular
in the fabrication of active origami.
DIW deposits viscous liquid inks in the form of a line onto
a platform.[360 ] The rheological properties of the ink are of crit-
ical importance, as shear-thinning and rapid transition to dila-
tant behavior are needed for efficient extrusion and subsequent
rapid shape-holding of the ink.[361 ] Because of this, rheological
modifiers, such as nanoclays, are often added to active mate-
rial inks, which can then be used to print origami structures
or hinges. The resolution depends on nozzle inner diameter
as well as the rheological properties of the ink. Since DIW uti-
lizes liquid resin, the deposited ink needs to be cured, either
directly after printing, or after the entire structure is printed.
Naficy et al.[ 354 ] developed a DIW printing method for hydrogels
that exhibit complex deformation by creating bilayer structures
with two different hydrogels, thermoresponsive pNIPAM and
thermally inactive poly(2-hydroxyethyl methacrylate) (pHEMA).
Using pHEMA as the base, active pNIPAM hinges were pat-
terned onto it, which were then surrounded by remaining inac-
tive material (Figure 9a-i). The folding box shown in Figure 9a-
ii experienced reversible folding and unfolding as the tempera-
ture changed below and above the pNIPAM’s LCST, respectively.
More complex deformation of hydrogel origami was demon-
strated by Gladman et al.,[362 ] who achieved biomimetic, plant-
like deformations of hydrogel bilayers upon immersion in wa-
ter. The ink used was mainly an acrylamide matrix embedded
with cellulose fibrils, which were aligned along the printing di-
rection during extrusion (Figure 9b-i, enhanced swelling along
the filament direction) and allowed for anisotropic swelling. Af-
ter immersion in water and swelling of the bilayers, complex
curvatures of flowers (Figure 9b-ii) were achieved, dictated by
the angles between the filaments of the bilayers. Aside from cel-
lulose, other materials such as nanomagnets have been used
in DIW printing of hydrogel. An acrylamide ink with NdFeB
magnets and nanoclay for shear-thinning was printed.[363] After
the matrix resin was cured, the magnetic hydrogel could be de-
formed and held in desired shapes for magnetization, imparting
a distributed magnetization within the material upon a large ap-
plied magnetic field. With this technique, origami folding such
as Miura–ori, as well as box and flower-like folding structures
were fabricated. Other works involving magnetic materials and
elastomers such as PDMS have demonstrated origami-like struc-
tures, such as a flapping butterfly[364] under a gradient mag-
netic field and reprogrammable movements of 3D magnetic soft
objects[365 ] via large impulse magnetic fields. First demonstrated
by Kim et al.,[37] DIW to program the magnetization of MSM inks
Adv. Mater. 2023, 2302066 © 2023 Wiley-VCH GmbH
2302066 (18 of 35)
15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202302066 by Stanford University, Wiley Online Library on [12/12/2023]. 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 www.advmat.de
Figure 9. Extrusion-based DIW and FFF for active origami structures. a) DIW of a hydrogel cross with thermoresponsive hydrogel hinges (Adapted with
permission.[354 ] Copyright 2016, Wiley). b) DIW of hydrogel with anisotropic cellulose ink for bilayer structures with complex bending (Adapted with
permission.[362 ] Copyright 2016, Springer Nature). c) Multimaterial DIW printing of MSM and magnetic SMP (Adapted with permission.[366 ] Copyright
2021, American Chemical Society). d) DIW of LCE with tunable actuation strain (Adapted with permission.[373 ] Copyright 2020, the Authors, some rights
reserved; exclusive licensee AAAS). e) Printing of LCE in 3D space onto supports which were then dissolved to result in a 3D LCE lattice (Adapted with
permission.[179 ] Copyright 2022, Wiley). f) FFF of SMP with programmed stress during printing (Adapted with permission.[ 378] Copyright 2017, The
Royal Society of Chemistry). g) FFF of cylindrical MSM embedded in a silicone matrix for biomimetic motions (Adapted with permission.[382 ] Copyright
2020, Elsevier).
along the print direction[37,366,367 ] allows for materials with pre-
cisely defined magnetization distributions (see also Figures 1c
and6b).Maetal.
[366 ] used this method in a multimaterial ap-
proach, such that MSMs and M-SMP were printed, enabling
metamaterials with segments of either M-SMP or MSM. The ob-
tained metamaterials had unique deformation modes, with the
MSMs deforming under magnetic fields and the M-SMPs requir-
ing heat for their deformation (Figure 9c). Other SMP origami via
DIW has been achieved, such as the work by Rodriguez et al.,[368 ]
which involved ink based on epoxidized soybean oil. With their
printing method, origami with initially flat or 3D shapes could be
achieved.
DIW is also well-suited for printing of LCEs.[167,369–372 ] Due to
the shear forces experienced by the ink during extrusion through
the nozzle, mesogens in LCE ink can be aligned along the print
direction,[369,370 ] thus programming the material. Roach et al.[180]
demonstrated the use of a room temperature printable LCE ink
for its use in actuating the hinges of multimaterial structures.
Adv. Mater. 2023, 2302066 © 2023 Wiley-VCH GmbH
2302066 (19 of 35)
15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202302066 by Stanford University, Wiley Online Library on [12/12/2023]. 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 www.advmat.de
Wang et al.[373] tuned printing parameters to show the variable
actuation strain of LCE filaments in folding structures. By vary-
ing printing parameters such as the printing temperature, the
nematic order of the inner filament could be tuned. For exam-
ple, with high printing temperatures, the inner part of a filament
largely remained in a polydomain state (Figure 9d-i), which re-
sulted in reduced actuation strain of the material after curing.
Using this approach, they demonstrated controllable actuated 3D
states of a flower-like substrate (Figure 9d-ii). DIW LCE printed
bilayers have also been demonstrated as hinges with controllable
fold angles[172 ] for robotics applications, as highlighted previously
in Figure 5h by Kotikian et al. Additionally, Peng et al.[179,374] have
developed fabrication strategies that combine DIW of LCEs with
other methods of 4D printing for multimaterial origami (as high-
lighted previously in Figure 5d). In a recent pioneering work, as
opposed to printing LCE onto a plate, they successfully extruded
LCE into 3D space, allowing for the creation of LCE-based struc-
tures that are not achievable by conventional DIW methods. As
shown in Figure 9e, by first printing dissolvable (non-LCE) struc-
tural supports and then using the support as a starting point of
an LCE fiber, LCE ink was extruded and then stretched to gener-
ate fibers in 3D space that maintained the ability of generating
high actuation strains. By printing numerous lines and then dis-
solving the structural supports, complex 3D LCE structures such
as the lattice pyramid in Figure 9e were achieved.
FFF (or FDM) involves the extrusion of melted thermoplas-
tic filaments, which experience solidification after cooling. The
main differences between DIW and FFF lie in material types: FFF
predominately prints thermoplastics and DIW is mainly used for
network polymers. SMPs (such as polyurethane or polyethylene)
printed by FDM have been used for programming of temporary
shapes.[137,375–377 ] In addition, tuning of printing parameters can
be used to introduce the desired stress during the printing pro-
cess, which later can be employed for shape change.[378–380 ] van
Manen et al.[ 378 ] utilized the polymer chain alignment along the
direction of extrusion enabled by the stretching of the melted
thermoplastic materials through the nozzle. The stretched de-
formation was memorized into the material after the filaments
cooled below Ttrans. By stacking layers with different print di-
rections (Figure 9f), deformations such as bending and twisting
were achievable, due to strain mismatch between layers upon
heating. Origami capable of water absorption has also been fab-
ricated by FFF. Baker et al.[381 ] printed hydrophilic polyurethane
sandwiched between two layers of hydrophobic polyurethane,
with gaps to create bilayer hinges than would bend when placed
in water. Lastly, FFF has also been used for magnetic material
origami,[382,383 ] in which FFF-printed magnetic components[382 ]
were oriented in a soft silicone matrix to program magnetiza-
tion and enable biomimetic motions (Figure 9g) as well as direct
printing of thermoplastic rubbers with incorporated soft mag-
netic particles.[383 ]
3.2.2. Digital Light Processing (DLP)
DLP is a vat-photopolymerization printing approach that involves
light projected onto a vat of photocurable resin, which cures
one layer of the resin at a time, after which the printing stage
is moved, allowing for subsequent layers of the desired struc-
ture to be printed. DLP is relatively fast, as an entire layer of
the resin can be cured at one time, with resolution possible at
tens of microns.[384,385 ] As DLP involves a vat of resin, it is often
challenging to achieve multimaterial printing, but Ge et al.[386 ]
addressed this by using an automated material exchange sys-
tem (Figure 10a), in which acrylate networks of different com-
positions could be used to create SMP structures, for example a
flower with petals of different Ttrans. Multimaterial DLP has also
been achieved by use of multicomponent resins in which mate-
rials could be preferentially cured under different wavelengths of
light, resulting in structures containing components of different
stiffness and swelling behaviors.[387 ] While SMPs often need to
be mechanically programmed after printing,[129] other methods
have avoided this by introducing differential crosslinking along
the thickness direction[388 ] to cause bending. Zhao et al.[389 ] did
this by utilizing frontal photopolymerization to fabricate Miura–
ori, a crane, and a variety of polyhedron. In a later work,[390] they
used DLP to introduce out-of-plane crosslinking gradients into
a flat film, which, upon immersion in water, resulted in desol-
vation of uncured chains from the film, inducing contraction to-
ward the side of the film that was less cured. By creating films
with fully cured panels and hinges that were differentially cross-
linked, origami structures were made which exhibited bending
at the hinges only (Figure 10b). The structures could be flat-
tened again by immersing the substrate in acetone, during which
the less cured region swelled. Another similar approach involv-
ing volatilization with an added post-curing step enabled origami
with enhanced material stiffness and the ability to hold apprecia-
ble weight.[391 ]
Another method to introduce different material properties
into the same network was achieved with the use of grayscale
light, such that light intensity varies within the projected light
pattern. With a two-stage curing approach, Kuang et al.[392] devel-
oped SMP structures with specific material properties imparted
both within the layers and throughout the part, involving, for
example, structures with materials of different moduli and Ttrans.
As shown in Figure 10c, a long strip with hinges of different
Ttrans allowed sequential folding of the SMP strip to a spiral
shape upon gradual heating. The advantage of this approach is
that it uses only one resin vat and retains DLP’s merits of high
speed and high resolution.
DLP has also been applied to LCEs with the assistance of post-
processing. Jin et al.[393 ] used a solvent-assisted programming ap-
proach. In their work, liquid crystalline organogels were used,
in which swollen and crosslinked 2D sheets or 3D LCE struc-
tures were deformed, followed by solvent evaporation in the de-
formed shape. The swelling of the organogels lead to flexibility
of the LC alignment and, upon deformation and subsequent dry-
ing, a new alignment of the mesogens was obtained, with mi-
cropores formed during drying helping to memorize the pro-
grammed director. Upon heating, the structure returned to its
printed shape (Figure 10d). Hydrogel origami fabrication has
also benefitted from DLP,[394–396 ] with a notable example from
Kim et al.,[394 ] in which biomimetic structures were achieved
with hydrogel bilayers made from silk fibroin bio-ink. The bi-
layer consisted of a flat layer and a patterned layer, where dif-
ferential swelling led to curling toward the patterned layer in
water. Structures such as a folding flower and venus flytrap
were demonstrated (Figure 10e). CANs have also been fabricated
Adv. Mater. 2023, 2302066 © 2023 Wiley-VCH GmbH
2302066 (20 of 35)
15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202302066 by Stanford University, Wiley Online Library on [12/12/2023]. 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 www.advmat.de
Figure 10. Digital light processing for fabrication of active origami. a) Multimaterial DLP of SMP origami structures (Reproduced with permissionf.[386]
Copyright 2016, Springer Nature). b) Differentially crosslinked and desolvation-induced folding of a Miura–ori sheet (Adapted with permission.[390]
Copyright 2017, Wiley). c) Grayscale DLP for SMP origami structures with tunable mechanical properties (Adapted with permission.[392 ] Copyright 2019,
the Authors, some rights reserved; exclusive licensee AAAS). d) Solvent-assisted programming and reprogramming of LCE structures (Adapted with
permission.[393 ] Copyright 2021, Wiley). e) Bioinspired silk hydrogel origami structures (Adapted with permission.[394 ] Copyright 2020, Elsevier). f)
Variety of foldable and remotely controllable DLP MSMs (Adapted with permission.[192] Copyright 2019, AAAS).
with DLP,[32,397,398 ] specifically, shape memory networks with self-
healing capability[398,399 ] and welding for modular assembly[32] of
origami structures. Although the use of photopolymerization via
DLP is typically not well-suited for the opacity of soft magnetic
materials, Xu et al.[192] utilized DLP to construct magnetic soft
robots with NdFeB particles that could be aligned to arbitrary 3D
magnetizations. Foldable structures such as an accordion sheet
and grippers (Figure 10f) were achieved, which could locomote
with cargo.
3.2.3. Inkjet Printing
Inkjet printing, which utilizes commercial printers such as
those from Stratasys, is a common method for the fabrica-
tion of multimaterial origami structures. Inkjet printers deposit
droplets of ink onto the printer bed, which is then smoothed
and photocured to produce layers of material (see Figure 11a).
Inkjet printers often have multiple printing heads, which en-
able printing with multiple materials, however, the materials
that can be used with inkjet printing are limited by those com-
mercially available, which are mainly SMP and hydrogels,[358]
along with the fact that these exact materials are not typi-
cally disclosed. Nevertheless, many origami works fabricated by
inkjet printing have been demonstrated, such as self-locking
(Figure 3e) or sequentially foldable (Figure 1c) boxes, Kresling-
based metamaterials (Figure 3h), pyramids (Figure 1c), flow-
ers (Figure 3f), and planes (Figure 5e) as shown in previ-
ous figures.[35,71,117,127,135,204,400,401 ] Another example is shown in
Figure 11b, where elastomer and glassy polymers are printed
Figure 11. Inkjet printing for active origami fabrication. a) Schematic of an inkjet printing process with multiple materials (Adapted with permission.[403]
Copyright 2017, AAAS). b) Rod-based multimaterial structure capable of folding from a flat lattice to 3D ball (Adapted with permission.[402 ] Copyright
2018, Elsevier).
Adv. Mater. 2023, 2302066 © 2023 Wiley-VCH GmbH
2302066 (21 of 35)
15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202302066 by Stanford University, Wiley Online Library on [12/12/2023]. 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 www.advmat.de
Figure 12. Micrometer and nanometer-scale fabrication techniques for active origami. a) Micrometer-scale molds used to fabricate different hydrogel
bilayer actuators which fold in water (Adapted with permission.[355 ] Copyright 2005, American Chemical Society). b) Direct laser writing/two-photon
polymerization to introduce voxels of different crosslinking densities into a hydrogel structure (Adapted with permission.[155] Copyright 2019, Elsevier).
c) Printing of liquid droplets to create bilayers of different osmolarities (Adapted with permission.[413 ] Copyright 2013, AAAS).
within the same rods to produce patterns that generate dif-
ferent types of deformation upon heating.[402 ] Notably, inkjet
printing has enabled the printing of “digital materials,[117,358,400 ]
which combine different ratios of base printing materials for
origami structures with different material properties and folding
capabilities.
3.3. Micro/Nano Scale Fabrication
As active origami shows great potential for applications in
biomedical environments as well as foldable electronic devices
such as microelectromechanical (MEMS) and nanoelectrome-
chanical systems (NEMS),[53] methods that can be used to
fabricate origami structures at the micrometer and nanome-
ter scale are very important.[404 ] There have been numer-
ous examples of micro and nano-scale origami largely com-
posed of metal.[334,405–408 ] The demonstrated methods to fold
these structures include surface tension of molten solder at
hinges[405,406 ] and chemical reactions.[334,407 ] However, these
methods are often not applicable to soft active materials. One
approach taken by Guan et al.[355] to achieve microscale fab-
rication of soft materials was the use of molds with microw-
ells to create bilayer hydrogel structures that bend in water
(Figure 12a).
4D printing at the nanoscale is also possible through di-
rect laser writing (DLW), which utilizes two-photon polymeriza-
tion (TPP) to fabricate structures at a resolution of as high as
100 nm.[6] Via TPP, a laser initiates two-photon absorption and
subsequent two-photon polymerization. The material near the
pulse of the laser polymerizes while the rest does not, with the
extent of polymerization depending on the power of the laser and
duration of pulse. Thus, high-resolution heterogeneous struc-
tures can be obtained. TPP has been used for fabrication of fold-
able voxelated LCE,[409,410 ] swellable materials,[155,411 ] and mag-
netic materials.[412 ] As shown in Figure 12b, Jin et al.[155 ] uti-
lized TPP for the tuned polymerization of voxels in a hydrogel
structure. By scanning the laser in the hydrogel precursor and
varying the laser power, voxels with different crosslinking den-
sity throughout the printed structure could be obtained. A pH-
responsive hydrogel umbrella structure with interspersed bilay-
ers was printed. Upon a change in pH, the hydrogel rim bi-
layer connected to the umbrella ribs would bend, causing rotation
of the ribs in a lever-like folding process. Tissue-like materials
can also be printed, as demonstrated by Villar et al.,[413] through
utilization of printed layers of droplets with different osmolari-
ties connected by lipid bilayers. When water flowed through the
bilayer, it caused the droplets to either swell or shrink, result-
ing in bending of the microscale structures toward the layer of
the initially more dilute droplets (Figure 12c). Aside from TPP,
other methods used in fabrication of active origami at microm-
eter and nanometer scales include photolithography[186,407,414,415 ]
and electron-beam lithography,[184,406] which use either light or
electron beams to pattern small features in the surface of a ma-
terial.
4. Origami Modeling
4.1. Modeling of Origami Structures
Structural modeling of origami enables the prediction of differ-
ent mechanisms’ deformations (single degree of freedom or mul-
tiple degrees of freedom) and mechanical behaviors (stiffness,
foldability, and stability). Under the guidance of structural mod-
eling, appropriately designed origami mechanisms permit pre-
ferred deformations and functions under mechanical loads or ex-
ternal stimuli. Various structural modeling approaches have been
developed by using bar and hinge models, shell-element-based
FEA models, and solid-element-based FEA models (Figure 13).
The bar and hinge model[416 ] is a simplified representation
of an origami structure that substitutes the creases between
panels with connected bars (green lines), and extra bars can
Adv. Mater. 2023, 2302066 © 2023 Wiley-VCH GmbH
2302066 (22 of 35)
15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202302066 by Stanford University, Wiley Online Library on [12/12/2023]. 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 www.advmat.de
Figure 13. Strategies for structural modeling of origami. a) Bar and hinge models of both conventional origami patterns (left, adapted with
permission.[417 ] Copyright 2013, American Physical Society) and curved-crease origami based on ideal torsional spring hinge without considering
hinge width (right, adapted with permission.[420 ] Copyright 2020, Elsevier). b) Bar and hinge model of origami with compliant hinges (adapted with
permission.[421 ] Copyright 2020, ASME). c) Shell-element-based finite element analysis (FEA) model with deformable panels connected by torsional
spring hinges (adapted with permission.[064 ] Copyright 2018, Elsevier). d) Shell-element-based finite element (FE) models of origami with compliant
hinges of an origami structure with smooth folds (adapted with permission.[426 ] Copyright 2016, Elsevier) and a bioinspired origami wing (adapted with
permission.[427 ] Copyright 2018, AAAS). e–h) Solid-element based FE models of origami with specially designed hinges with e) reduced hinge thickness
(adapted with permission.[428 ] Copyright 2018, The Royal Society of Chemistry.), f) introduced cuts at hinges (Adapted with permission.[075 ] Copyright
2022, the Authors, some rights reserved; exclusive licensee AAAS.), g) reduced material stiffness (Top: adapted with permission.[071] Copyright 2020,
Wiley. Bottom: adapted with permission.[135] Copyright 2017, Springer Nature.) and h) multilayer structure (adapted with permission.[125 ] Copyright
2019, National Academy of Sciences).
be added across the origami panel (blue line) to accommo-
date for the in-plane stiffness (stretching and shearing), as il-
lustrated in the left of Figure 13a.[417] Torsional hinge springs
are placed along the bars between panels and along the bars
across panels to model origami hinge folding and out-of-plane
panel bending, respectively. Various types of bar and hinge
models with one,[3,68,416,417 ] two,[418 ] or four[ 419] bars across the
panel have been developed. These structural modeling meth-
ods were applied to predict both linear and nonlinear responses
of traditional origami patterns (Figure 13a, left),[417] curved-
creased origami (Figure 13a, right),[420] as well as origami struc-
tures accounting for compliant hinges with a finite width be-
tween the panels (Figure 13b)[421] and active hinges with stimuli-
responsive actuations.[422 ] Based on the bar and hinge mod-
els, software such as Origami Simulator,[423 ] MERLIN2,[424 ]
and Tessellatica[425 ] have been developed for origami structural
modeling.
Although the bar and hinge models help reduce the cost of cal-
culations, it leaves out valuable information such as stress and
strain distributions in the origami structure during the folding
process. FEA is a widely used simulation method based on phys-
ical modeling, providing a higher accuracy compared to the bar
and hinge model. The shell-element-based FEA model is straight-
forward to adopt, as the thickness of origami panels is gener-
ally much smaller than the in-plane dimension.[61,64,98,127 ] The
shell-element-based origami panels consider not only in-plane
stretching and shearing, but out-of-plane bending as well. To
account for the rotation of panels about the hinges, both ideal
rotational spring hinges with zero width (Figure 13c)[64] and com-
pliant hinges with finite width (Figure 13d, left[426] and right[ 427 ])
Adv. Mater. 2023, 2302066 © 2023 Wiley-VCH GmbH
2302066 (23 of 35)
15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202302066 by Stanford University, Wiley Online Library on [12/12/2023]. 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 www.advmat.de
have been used, which are capable of capturing highly nonlinear
mechanical responses such as snap-through deformations.[64,427 ]
FEA simulations of origami structures based on 3D solid ele-
ments consider the finite thickness of origami panels and hinges.
Although this simulation technique is more computationally ex-
pensive than the others, the 3D modeling can take into account
the complex geometry of origami, for example, layered origami
structures composed of active and inactive materials, and thus
provides the most information for the origami folding process.
In addition, other physical information such as temperature gra-
dient, swelling due to water absorption, etc. can be incorpo-
rated. Considering the actual hinge width and thickness in the
3D models, different methods have been adopted to ensure that
the folding of panels happens at preferred positions, namely at
the compliant hinges (Figure 13e–h).[71,75,125,135,428] Reduction in
hinge thickness (Figure 13e)[428] or introduction of cuts at hinges
(Figure 13f)[75] are generally used to generate compliant hinges
in structures based on a single material.[307,429 ] When multima-
terial fabrication is allowed, reduction in the material stiffness
(Figure 13g, top[71] and bottom[135 ] ) at hinges is an alternative
option to reduce the effective hinge stiffness. As discussed be-
fore, layered structures combining active materials and inactive
materials are commonly used for shape reconfiguration of ac-
tive origami. Correspondingly, 3D FEA models with multilayer
structures (Figure 13h)[125] are also widely seen in different works
for the prediction of active origami actuation.[30,101,135,138,161,430 ]
It should be noted that the dynamics of actuation become im-
portant to the folding behaviors of origami structures, especially
for soft materials which can exhibit instabilities when actuated
rapidly.
4.2. Modeling of Active Materials
Active origami modeling inevitably requires modeling of the me-
chanical deformation and stresses of constituent active materials
in response to external stimuli. Depending on the material type,
the mechanical behaviors are coupled with certain physical fields.
Continuum mechanics-based constitutive models typically give
the equations describing how state functions, such as internal
energy, entropy, and stress, depend on state variables such as de-
formation gradient (or strain) and temperature, and sometimes
on internal variables specific to each physical field. Incorporation
of constitutive models with numerical simulation tools, such as
FEA, enables the simulation of active response of materials and
origami under mechanical and multiphysical stimuli.
Active materials used in origami are often based on polymeric
materials, which exhibit large deformations. In this case, the
multiphysics behaviors (or constitutive behaviors) are generally
described by first specifying the Helmholtz free energy as a func-
tion of macroscopic deformation and internal variables associ-
ated with the coupled physical fields.[431 ] The stress–strain rela-
tionship can then be given by taking the derivative of the free
energy with respect to the deformation. For instance, the Cauchy
stress 𝝈is given by:
𝝈=J1𝜕
𝜕FFT(1)
where Ψis the Helmholtz free energy density, Fis the deforma-
tion gradient, the superscript “T” represents the transpose, and J
=det(F) is the volume change. For a purely elastic network, the
free energy, denoted by Ψstretch, recovers the strain energy, which
depends on the configurational entropy due to network stretch-
ing only. Some specific forms of Ψstretch are available, such as neo-
Hookean,[432 ] Arruda–Boyce,[433 ] Ogden,[434 ] etc. Note that, in ad-
dition to Equation 1, certain equations that constrain other state
or internal variables may need to be satisfied due to the laws of
thermodynamics. A summary of the modeling of different active
materials is given in Table 3.
4.2.1. SMPs
A large number of models have been developed for SMPs, as
summarized in the literature.[196,471 ] These models can be divided
into two classes: thermoviscoelastic models[197,435,436,439,440,443 ]
and phase evolution models.[437,438,441,442 ] The thermoviscoelas-
tic model has been widely used for amorphous polymers, in
which the shape memory effects are due to the change in chain
mobility during glass transition. Based on this mechanism, the
general idea is to adopt the temperature-dependent viscosity in
the framework of viscoelasticity, thus leading to the term “ther-
moviscoelasticity”. For example, assuming the relaxation time
to follow the time-temperature superposition principle (TTSP)
in simple viscoelastic models, the shape memory effects can be
captured.[436,443 ] Multibranch viscoelastic models are often used
to accommodate multiple relaxation modes in real polymers. In-
tegration of TTSP with multibranch models showed improved
capability in predicting the shape memory behaviors[197,435,440 ] or
capturing multi-shape memory effects.[439 ]
The phase evolution model captures the physics in semi-
crystalline polymers, although it is also used to model amorphous
SMPs, in a so-called phenomenological manner. In the phase
evolution model, the material is continuously crystallized in a
loading process and the newly formed polymer crystals are as-
sumed stress-free. Crystals formed at different instants can be
seen as different phases, which undergo different deformation
histories and thus carry different free energies. With this concept
as the foundation, the Helmholtz free energy density function
can be expressed as:[464 ]
Ψ=CORΨstretch Fe
0t+t
0
Δ(𝜏,t
)Ψstretch Fe
𝜏td𝜏(2)
where COR and Δ(𝜏,t)d𝜏are the fractions of the originally ex-
isted phases and the later formed phases, respectively, Feis the
mechanical deformation gradient for different phases with the
subscript t1t2(or 0t) denoting the deformation from time t1
to t2(or 0 to t), and Ψstretch is the free energy function due to net-
work stretching. Note that COR and Δ(𝜏,t) evolve based on certain
kinetic laws[441 ] to capture the shape memory behaviors. All the
evolving fractions in Equation 2 need to be tracked in numerical
simulations, which is computationally expensive. The effective
phase model has been developed to address this issue.[472 ] In ad-
dition, the concept of phase evolution has been used not only in
the shape memory effects of semi-crystalline polymers but also
in the shape memory or viscoelastic behavior of many active poly-
mers, such as CANs. This will be discussed more in Section 4.2.5.
Adv. Mater. 2023, 2302066 © 2023 Wiley-VCH GmbH
2302066 (24 of 35)
15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202302066 by Stanford University, Wiley Online Library on [12/12/2023]. 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 www.advmat.de
Table 3. Summary of modeling of active materials.
Material Additional terms in free energy State variables and internal variables
SMPs[197,435–443 ] Deformations;
phase fractions;
temperature; light; moisture
Hydrogels[444–449 ] Free energy due to substance mixing Deformations;
chemical species;
temperature
LCEs[252,450–457 ] Nematic free energy Deformations;
mesogen orientations;temperature; light
MSMs[356,458–463 ] Magnetic potential Deformations;
magnetic field; magnetization density
CANs[464–470 ] Deformations;
phase fractions;
temperature; light; moisture
4.2.2. Hydrogels
Hydrogels can exhibit reversible volumetric changes by absorb-
ing or repelling water (or solvent). Thermodynamics and me-
chanics theories have been developed to capture the physical cou-
pling between the stretching of the network, mixing of the solvent
and the polymer, and network swelling.[444,446,447] In the pioneer-
ingworkbyHongetal.,
[444 ] the Helmholtz free energy density
was expressed in two parts,
Ψ=Ψ
stretch
mix (3)
where Ψstretch and Ψmix represent the free energy due to the
stretching and the mixing, respectively. A commonly usedΨstretch
based on the Gaussian chain distribution is:[473 ]
Ψstretch =1
2NkT trace FTF32lnJ(4)
where Nis the number of crosslinks per unit reference volume,
kis Boltzmann constant, and Tis the temperature. Note that
Equation 4 is slightly different from the neo-Hookean form. Non-
Gaussian statistics-based forms of Ψstretch have also been used to
account for the limited extensibility of polymer chains.[447,448 ] The
free energy Ψmix due to the mixing of polymer and the solvent is
written based on the Flory–Huggins model:[474,475]
Ψmix =kTC ln vC
1+vC +𝜒
1+vC (5)
where vis the volume of a fluid molecule, Cis the number of fluid
molecules per unit reference volume, and 𝜒is the Flory–Huggins
interaction parameter characterizing the disaffinity between the
polymer and the fluid.
With the above free energy form, Hong et al.[444] derived the
constitutive relations for the stress, chemical potential and fluid
flux under isothermal conditions. Later, Chester et al.[447,448] and
Duda et al.[446 ] developed general, continuum-mechanics frame-
works that could describe behaviors under complex thermo-
chemo-mechanical loadings. The temperature-dependent 𝜒(T)
has also been used to model hydrogels with temperature-driven
swelling/deswelling properties, such as pNIPAM.[448,449 ]
Implementation of the constitutive relations of hydrogels into
FEA codes enables the direct simulation of hydrogels with com-
plex geometry and under complex loadings.[448,476,477 ] For imple-
mentations in the FEA package ABAQUS (Dassault Systèmes,
France), the user hyperelastic material (UHYPER)[476] or user el-
ement (UEL)[248,252,478 ] the mesogens may be reoriented and LCE
may experience a phase transition between an ordered (nematic)
and disordered (isotropic) state. To describe the intrinsic cou-
pling between the deformation, reorientation of nematic order
and other stimuli fields, many constitutive models have been
developed.[252,450–453 ] Generally, the Helmholtz free energy den-
sity can be expressed as:[452 ]
Ψ=Ψ
stretch
nematic (6)
where Ψstretch is the free energy due to network stretching
and Ψnematic is the nematic free energy.[248,452] The Ψstretch form
first given by Bladon et al.[479 ] takes the energy form of neo-
Hookean but is measured between the supposed isotropic phase
and the current configuration, which has been widely used in
literature.[450–454 ] The Ψnematic can be written as a function of
the nematic order, and various forms have been used.[450,452,480]
For polydomain LCEs, the free energy should take different
forms in different nematic regions.[451,453 ] In general, the to-
tal free energy is a function of the network stretching, pre-
ferred orientation, and the nematic order. Recently, constitutive
models have been developed for heat-[452 ] or light-responsive[455]
LCEs, as well as LCEs with internal dissipations.[456,457 ] In
addition, in many applications, the anisotropic thermal ex-
pansion of LCE with experimentally measured coefficients
has been widely used in FEA to mimic the actuation of
LCEs.[179,208,373 ]
4.2.3. MSMs
MSMs can be classified into soft-magnetic and hard-magnetic
materials depending on the coercivity of the embedded mag-
netic particles. Many constitutive models on soft-magnetic elas-
tomers have been developed.[481–483 ] However, here the hard-
MSMs which are capable of complex active shape changes
Adv. Mater. 2023, 2302066 © 2023 Wiley-VCH GmbH
2302066 (25 of 35)
15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202302066 by Stanford University, Wiley Online Library on [12/12/2023]. 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 www.advmat.de
will be focused on. Zhao et al.[458 ] developed the first gen-
eral continuum framework for 3D finite deformation of hard-
MSMs, in which the Helmholtz free energy density consists of
two parts:
Ψ=Ψ
stretch
magnetic (7)
where Ψstretch represents the free energy due to the stretching of
the soft matrix and the magnetic-related free energy Ψmagnetic can
be expressed as the magnetic potential of the embedded hard-
magnetic particles, that is,
Ψmagnetic =−FM Bapplied (8)
where Mis the referential magnetization density and Bapplied is
the applied magnetic field. Note that with the above Helmholtz
free energy, Zhao et al.[458] derived the constitutive relation for
the stress of hard-MSMs, which has been implemented into
ABAQUS through the user element (UEL) subroutine. Since the
work of Zhao et al.,[458 ] many reduced-order models for hard-
MSMs have also been developed to enable computationally low-
cost simulations and designs.[460–462,484,485 ]
4.2.4. CANs
Modeling macroscopic behaviors of CANs generally requires in-
corporating the molecular-level kinetics of chemical reactions
into the framework of continuum-level constitutive theory. Ac-
cording to how they are coupled, three main types of models
which utilize different strategies have been developed. The first
model is based on the concept of phase evolution. Long et al.[466 ]
developed the first constitutive model for a light-activated asso-
ciative CAN network based on this concept. Since then, the same
concept has been applied to CAN systems with various types
of stimuli and dynamic bonds.[464–466,472,486 ] Here, the phase is
a more general representation of a small collection of polymer
chains. Upon external stimuli, the polymer chains can contin-
uously evolve due to the bond cleavage/re-formation. Using a
similar concept for the crystallization, here the newly reformed
chains are assumed stress-free. Therefore, the Helmholtz free
energy density can be given by Equation 2.[464 ] Kinetic laws
for the phase evolution are governed by the kinetics of chem-
ical reactions and thus coupled to certain stimulus fields (e.g.,
temperature, light, and chemical species). The phase evolution
model has been integrated with FEA, enabling predictive sim-
ulations of viscoelastic behaviors of CANs with complex 3D
geometries.[487–490 ]
The second model was developed by Vernerey et al.,[468,469] who
expressed the total free energy density in terms of a statistical
distribution of the chain end-to-end vector r, that is,
Ψ=Ω
𝜙(r,t
)𝜓C(r)dΩ(9)
where ϕis the end-to-end vector distribution, 𝜓Cis the free en-
ergy density of the single chain, dΩ=r2sin𝜃d𝜃d𝜔is a small vol-
ume element of the conformation space Ω. Note that ϕ(r,t)de-
scribes the density of chains whose ris within the element dΩ.
Upon deformation and chemical reactions, the distribution func-
tion is evolved following certain kinetic laws associated with the
reaction kinetics, which can capture the macroscopic CAN behav-
iors.
The above two models physically incorporate the molecular-
level reaction kinetics, as discussed in two recent reviews.[491,492 ]
In addition, the third type of approach incorporates effects of dy-
namic reactions into phenomenological viscoelastic or viscoplas-
tic frameworks.[470,493 ] Thegeneralideaistousearheological
model with certain combinations of springs and dashpots to de-
scribe the CAN behavior. The viscous flow behavior of different
dashpots can be chosen to depend on different stimuli follow-
ing the reaction kinetics or the TTSP to capture complex multi-
physics behaviors.
5. Conclusion
Active origami structures are appealing for many applications
in reconfigurable devices, robotics, biomimetic actuators, meta-
materials, aerospace, and biomedical environments and have
seen rapid development in recent years. In this review, we
have discussed common active materials and their actuation
mechanisms, along with applications of active origami, com-
mon origami fabrication techniques, and a brief overview of the
modeling of active materials and origami structures. Although
we have highlighted here the many exciting advances in the
field of active origami, such as structures with reprogrammable
deformation, locomotive multifunctional origami robots, and
state-of-the-art fabrication techniques to enable self-folding,
we would like to point out the primary challenges currently
faced, as well as the pertinent opportunities related to active
origami.
While the low modulus of active polymers allows these ma-
terials to be applicable for several biomedical applications, soft
polymer-based structures also suffer from low actuation force
and poor mechanical strength, which is a significant limitation
for many practical applications. Deeper exploration into spe-
cial design of origami hinges[493 ] as well as further develop-
ment of composite active materials[205,494–496 ] or methods to store
elastic potential energy[215 ] may be able to help address these
limitations.
Within the realm of fabrication, multimaterial structures,
which can be used in multi-stimuli response applications and
enable complex structures with distributed material rigidity, still
face limitations such as the compatibility of materials, as well
as printers capable of handling multiple different materials or
even different printing methods within one setup.[179,374 ] Devel-
opment of printers that could achieve this at a relatively low cost
remains a challenge. Additionally, while rapid 4D printing meth-
ods such as DLP have been well developed for SMPs, further de-
velopment of similar rapid 4D printing methods for materials
such as hydrogels, LCEs, and MSMs would be beneficial. In ad-
dition to small-scale fabrication, the commercialization or wide-
scale production of these materials becomes important. Hydro-
gels have been commercialized for use in several everyday items
such as wound dressings and contact lenses[498 ] while CANs have
been used for material recycling and reprocessing. Certain kinds
of shape memory polymers are popular as heat shrink tubes or
shrinkable plastic sheets while Nitinol, a type of shape memory
Adv. Mater. 2023, 2302066 © 2023 Wiley-VCH GmbH
2302066 (26 of 35)
15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202302066 by Stanford University, Wiley Online Library on [12/12/2023]. 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 www.advmat.de
alloy, has been widely applied within biomedical, automotive,and
aerospace industries. For a material with complicated fabrication
such as LCE, however, industrialization may take considerably
longer.
In addition, the design of accurate, controllable folding mech-
anisms is of great importance for practical applications of
complex foldable structures. In terms of minimizing materials
and weight of a structure, a simplified actuator arrangement is
preferred that can still result in precise folding. The exploration
of instabilities[499 ] and buckling mechanisms to enable precise
folding and greater actuation force is suggested for deeper study,
to address the limitations currently seen in the folding of soft ma-
terials.
Important to the future of active origami are systems which
can effectively integrate sensing capabilities for data acquisi-
tion in a variety of applications. While the ability to change
states in response to stimulus is inherent to the design of ac-
tive origami systems, those which can adapt to environments
and self-sense could result in nearly autonomous machines,
especially for robotics and soft actuator applications. As ac-
tive origami has enabled many biomimetic soft robotics ap-
plications, to further emulate these biological systems, more
sophisticated interactions between the soft robot and its en-
vironment are required. There have been recent works which
take advantage of active origami sensing,[73,220,500,501 ] however,
the demonstrated sensing capabilities are still rather undevel-
oped and should be studied further. Ideally, these origami sys-
tems would be tetherless or have on-board control systems ca-
pable of intaking sensory information from the surrounding
environment.
Overall, the field of active origami has seen many exciting ad-
vancements in the past few years, which can be used toward
self-deployable and foldable structures on multiple size scales for
many aerospace, metamaterial, robotic, and biomedical applica-
tions. With the further development and exploration of structural
and material designs for active origami, unique solutions can be
reached for many engineering challenges seen in the abovemen-
tioned areas.
Acknowledgements
R.R.Z., S.L., and S.W. acknowledge support from the NSF Award CPS-
2201344 and the NSF Career Award CMMI-2145601. H.J.Q. and X.S. ac-
knowledge the support of an AFOSR grant (FA9550-20-1-0306; Dr. B.-L.
“Les” Lee, Program Manager).
Conflict of Interest
The authors declare no conflict of interest.
Keywords
active materials, folding, origami, origami modeling, stimuli-responsive
materials
Received: March 4, 2023
Revised: April 13, 2023
Published online:
[1] T. Chen, O. R. Bilal, R. Lang, C. Daraio, K. Shea, Phys. Rev. Appl. 2019,
11, 064069.
[2] M. Schenk, A. D. Viquerat, K. A. Seffen, S. D. Guest, J. Spacecr. Rock-
ets 2014,51, 762.
[3] J. L. Silverberg, A. A. Evans, L. Mcleod, R. C. Hayward, T. Hull, C. D.
Santangelo, I. Cohen, Science 2014,345, 647.
[4] J. T. B. Overvelde, T. A. De Jong, Y. Shevchenko, S. A. Becerra, G. M.
Whitesides, J. C. Weaver, C. Hoberman, K. Bertoldi, Nat. Commun.
2016,7, 10929.
[5] E.T.Filipov,T.Tachi,G.H.Paulino,Proc. Natl. Acad. Sci. USA 2015,
112, 12321.
[6] Z. Lin, L. S. Novelino, H. Wei, N. A. Alderete, G. H. Paulino, H. D.
Espinosa, S. Krishnaswamy, Small 2020,16, 2002229.
[7] Z. Zhai, L. Wu, H. Jiang, Appl. Phys. Rev. 2021,8, 041319.
[8] A. Pagano, T. Yan, B. Chien, A. Wissa, S. Tawfick, Smart Mater. Struct.
2017,26, 094007.
[9] S.-J. Kim, D.-Y. Lee, G.-P. Jung, K.-J. Cho, Sci. Rob. 2018,3, eaar2915.
[10] S. Felton, M. Tolley, E. Demaine, D. Rus, R. Wood, Science 2014,345,
644.
[11] R. V. Martinez, C. R. Fish, X. Chen, G. M. Whitesides, Adv. Funct.
Mater. 2012,22, 1376.
[12] K. Kuribayashi, K. Tsuchiya, Z. You, D. Tomus, M. Umemoto, T. Ito,
M. Sasaki, Mater. Sci. Eng., A 2006,419, 131.
[13] S. Miyashita, S. Guitron, K. Yoshida, S. Li, D. D. Damian, D. Rus, in
2016 IEEE Int. Conf. on Robot. and Autom. (ICRA), Stockholm 2016.
[14] Q. Cheng, Z. Song, T. Ma, B. B. Smith, R. Tang, H. Yu, H. Jiang, C.
K. Chan, Nano Lett. 2013,13, 4969.
[15] S. A. Nauroze, L. S. Novelino, M. M. Tentzeris, G. H. Paulino, Proc.
Natl. Acad. Sci. USA 2018,115, 13210.
[16] Z. Song, T. Ma, R. Tang, Q. Cheng, Xu Wang, D. Krishnaraju, R.
Panat, C. K. Chan, H. Yu, H. Jiang, Nat. Commun. 2014,5, 3140.
[17] G. Masera, M. Pesenti, F. Fiorito, J. Archit. Eng. 2018,24, 04018022.
[18] S. A. Zirbel, R. J. Lang, M. W. Thomson, D. A. Sigel, P. E. Walkemeyer,
B. P. Trease, S. P. Magleby, L. L. Howell, J. Mech. Des. 2013,135,
111005.
[19] S. A. Zirbel, B. P. Trease, M. W. Thomson, R. J. Lang, S. P. Magleby,
L. H. Howell, Micro- and Nanotechnology Sensors, Systems, and Ap-
plications VII,2015, 94671C.
[20] J. Morgan, S. P. Magleby, L. L. Howell, J. Mech. Des. 2016,138,
052301.
[21] J. Butler, J. Morgan, N. Pehrson, K. Tolman, T. Bateman, S. P.
Magleby, L. L. Howell, in Int. Des. Eng. Tech. Conf. and Comput. and
Inform. in Eng. Conf., ASME, Charlotte, North Carolina, USA, 2016.
[22] B. J. Edmondson, L. A. Bowen, C. L. Grames, S. P. Magleby, L. L.
Howell, T. C. Bateman, in ASME 2013 Conf. on Smart Mater., Adap-
tive Struct. and Intell. Syst., Snowbird, Utah, USA 2013.
[23] H. Suzuki, R. J. Wood, Nat. Mach. Intell. 2020,2, 437.
[24] P. Bhovad, J. Kaufmann, S. Li, Extreme. Mech. Lett. 2019,32, 100552.
[25] H. Yasuda, T. Tachi, M. Lee, J. Yang, Nat. Commun. 2017,8, 962.
[26] T. Liu, Y. Wang, K. Lee, IEEE Robot Autom. Lett. 2017,3, 116.
[27] D. Melancon, B. Gorissen, C. J. García-Mora, C. Hoberman, K.
Bertoldi, Nature 2021,592, 545.
[28] S.Li,D.M.Vogt,D.Rus,R.J.Wood,Proc. Natl. Acad. Sci. USA 2017,
114, 13132.
[29] W. Kim, J. Byun, J.-K. Kim, W.-Y. Choi, K. Jakobsen, J. Jakobsen, D.-Y.
Lee, K.-J. Cho, Sci. Rob. 2019,4, eaay3493.
[30] C. Yoon, R. Xiao, J. Park, J. Cha, T. D. Nguyen, D. H. Gracias, Smart
Mater. Struct. 2014,23, 094008.
[31] Q. Ze, S. Wu, J. Dai, S. Leanza, G. Ikeda, P. C. Yang, G. Iaccarino, R.
R. Zhao, Nat. Commun. 2022,13, 3118.
[32] Z. Fang, H. Song, Y. Zhang, B. Jin, J. Wu, Q. Zhao, T. Xie, Matter
2020,2, 1187.
[33] C.Yuan,D.J.Roach,C.K.Dunn,Q.Mu,X.Kuang,C.M.Yakacki,T.
J. Wang, K. Yu, H. J. Qi, Soft Matter 2017,13, 5558.
Adv. Mater. 2023, 2302066 © 2023 Wiley-VCH GmbH
2302066 (27 of 35)
15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202302066 by Stanford University, Wiley Online Library on [12/12/2023]. 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 www.advmat.de
[34] J. Ryu, M. D’amato, X. Cui, K. N. Long, H. J. Qi, M. L. Dunn, Appl.
Phys. Lett. 2012,100, 161908.
[35] Q. Ge, C. K. Dunn, H. J. Qi, M. L. Dunn, Smart Mater. Struct. 2014,
23, 094007.
[36] E. A. Peraza-Hernandez, D. J. Hartl, R. J. Malak Jr, D. C. Lagoudas,
Smart Mater. Struct. 2014,23, 094001.
[37] Y.Kim,H.Yuk,R.Zhao,S.A.Chester,X.Zhao,Nature 2018,558,
274.
[38] Y. Kim, X. Zhao, Chem. Rev. 2022,122, 5317.
[39] Y. Park, T. S. Chung, G. Lee, J. A. Rogers, Chem. Rev. 2022,122, 5277.
[40] J. M. Mccracken, B. R. Donovan, T. J. White, Adv. Mater. 2020,32,
1906564.
[41] H. Yuk, J. Wu, X. Zhao, Nat. Rev. Mater. 2022,7, 935.
[42] I. Apsite, S. Salehi, L. Ionov, Chem. Rev. 2022,122, 1349.
[43] A. Kirillova, L. Ionov, J. Mater. Chem. B 2019,7, 1597.
[44] S. Wu, W. Hu, Q. Ze, M. Sitti, R. Zhao, Multifunct. Mater. 2020,3,
042003.
[45] X. Ning, X. Wang, Y. Zhang, X. Yu, D. Choi, N. Zheng, D. S. Kim,
Y. Huang, Y. Zhang, J. A. Rogers, Adv. Mater. Interfaces 2018,5,
1800284.
[46] D. Rus, M. T. Tolley, Nat. Rev. Mater. 2018,3, 101.
[47] S. Li, H. Fang, S. Sadeghi, P. Bhovad, K.-W. Wang, Adv. Mater. 2019,
31, 1805282.
[48] V. A. Bolaños Quiñones, H. Zhu, A. A. Solovev, Y. Mei, D. H. Gracias,
Adv. Biosyst. 2018,2, 1800230.
[49] Y. Zhu, M. Schenk, E. T. Filipov, Appl. Mech. Rev. 2022,74, 030801.
[50] J. J. Park, P. Won, S. H. Ko, Int. J. Precis. Eng. Manuf. Green Technol.
2019,6.
[51] L. Xu, T. C. Shyu, N. A. Kotov, ACS Nano 2017,11, 7587.
[52] S. J. P. Callens, A. A. Zadpoor, Mater. Today 2018,21, 241.
[53] J.Rogers,Y.Huang,O.G.Schmidt,D.H.Gracias,MRS Bull. 2016,
41, 123.
[54] K. Miura, The Institute of Space and Astronautical Sci. Rep.1985,618,
p. 1.
[55] D.-Y. Lee, J.-S. Kim, S.-R. Kim, J.-S. Koh, K.-J. Cho, in Int. Des. Eng.
Tech. Conf. and Comput. and Inform. in Eng. Conf.,AmericanSociety
of Mechanical Engineers, New York 2013.
[56] S.Li,J.J.Stampi,H.J.Xu,E.Malkin,E.V.Diaz,D.Rus,R.J.
Wood, in 2019 Int. Conf. on Robot. and Autom. (ICRA), Montreal,
QC, Canada 2019.
[57] H. Yasuda, T. Yein, T. Tachi, K. Miura, M. Taya, Proc. R. Soc. A 2013,
469, 20130351.
[58] M. Schenk, S. D. Guest, Proc. Natl. Acad. Sci. USA 2013,110, 3276.
[59] J. L. Silverberg, J.-H. Na, A. A. Evans, B. Liu, T. C. Hull, C. D.
Santangelo, R. J. Lang, R. C. Hayward, I. Cohen, Nat. Mater. 2015,
14, 389.
[60] Y. Yoshimura, No. NACA-TM-1390., 1955, 93R23165.
[61] C. Lv, D. Krishnaraju, G. Konjevod, H. Yu, H. Jiang, Sci. Rep. 2014,
4, 5979.
[62] T. Tachi, J. Mech. Des. 2013,135, 111006.
[63] B. Kresling, in Origami3(Ed: T. Hull), Taylor & Francis Group, Cali-
fornia, 2002, Part II, p. 197.
[64] N.Nayakanti,S.H.Tawck,A.J.Hart,Extreme Mech. Lett. 2018,21,
17.
[65] Z. Zhai, Y. Wang, H. Jiang, Proc. Natl. Acad. Sci. USA 2018,115, 2032.
[66] K. Liu, T. Tachi, G. H. Paulino, Nat. Commun. 2019,10, 4238.
[67] E. T. Filipov, M. Redoutey, Extreme Mech. Lett. 2018,25, 16.
[68] L. H. Dudte, E. Vouga, T. Tachi, L. Mahadevan, Nat. Mater. 2016,15,
583.
[69] J. Kaufmann, S. Li, Soft Robot. 2021, 9, 212.
[70] Lu Lu, X. Dang, F. Feng, P. Lv, H. Duan, Proc. R. Soc. A 2022,478,
20210712.
[71] L.-C. Wang, W.-L. Song, Y.-J. Zhang, M.-J. Qu, Z. Zhao, M. Chen, Y.
Yang, H. Chen, D. Fang, Adv. Funct. Mater. 2020,30, 1909087.
[72] S. Sengupta, S. Li, J. Intell. Mater Syst. Struct. 2018,29, 2933.
[73] L. S. Novelino, Q. Ze, S. Wu, G. H. Paulino, R. Zhao, Proc. Natl.
Acad. Sci. USA 2020,117, 24096.
[74] Z. Li, N. Kidambi, L. Wang, K.-W. Wang, Extreme Mech. Lett. 2020,
39, 100795.
[75] Q. Ze, S. Wu, J. Nishikawa, J. Dai, Y. Sun, S. Leanza, C. Zemelka, L.
S. Novelino, G. H. Paulino, R. R. Zhao, Sci. Adv. 2022,8, eabm7834.
[76] B. Treml, A. Gillman, P. Buskohl, R. Vaia, Proc. Natl. Acad. Sci. USA
2018,115, 6916.
[77] S. Xu, Z. Yan, K.-I. Jang, W. Huang, H. Fu, J. Kim, Z. Wei, M. Flavin,
J. Mccracken, R. Wang, A. Badea, Y. Liu, D. Xiao, G. Zhou, J. Lee, H.
U. Chung, H. Cheng, W. Ren, A. Banks, X. Li, U. Paik, R. G. Nuzzo,
Y. Huang, Y. Zhang, J. A. Rogers, Science 2015,347, 154.
[78] N. A. Alderete, L. Medina, L. Lamberti, C. Sciammarella, H. D.
Espinosa, Extreme Mech. Lett. 2021,43, 101146.
[79] X. Zhang, L. Medina, H. Cai, V. Aksyuk, H. D. Espinosa, D. Lopez,
Adv. Mater 2021,33, 2005275.
[80] V. G. A. Goss, G. H. M. Van Der Heijden, J. M. T. Thompson, S.
Neukirch, Exp. Mech. 2005,45, 101.
[81] X. Lachenal, P.M. Weaver, S. Daynes, Proc.R.Soc.A2012,468,1230.
[82] Y. Goto, Y. Watanabe, T. Kasugai, M. Obata, Int. J. Solids Struct. 1992,
29, 893.
[83] P.-O. Mouthuy, M. Coulombier, T. Pardoen, J.-P. Raskin, A. M. Jonas,
Nat. Commun. 2012,3, 1290.
[84] P. F. Pai, A. N. Palazotto, Int. J. Solids Struct. 1996,33, 1335.
[85] B. Audoly, K. A. Seffen, J. Elast. 2015,119, 293.
[86] S.Wu,L.Yue,Y.Jin,X.Sun,C.Zemelka,H.J.Qi,R.Zhao,Adv. Intell.
Syst. 2021,3, 2100107.
[87] X. Sun, S. Wu, J. Dai, S. Leanza, L. Yue, L. Yu, Y. Jin, H. J. Qi, R. R.
Zhao, Int. J. Solids Struct. 2022,248, 111685.
[88] S. Leanza, S. Wu, J. Dai, R. R. Zhao, J. Appl. Mech. 2022,89.
[89] L. Lu, S. Leanza, J. Dai, X. Sun, R. R. Zhao, J. Mech. Phys. Solids 2022,
171, 105142.
[90] S. Wu, J. Dai, S. Leanza, R. R. Zhao, Extreme Mech. Lett. 2022,53,
101713.
[91] E. D. Demaine, M. L. Demaine, D. Koschitz, T. Tachi,Symmetry: Cult.
Sci. 2015,26, 145.
[92] M. A. Dias, L. H. Dudte, L. Mahadevan, C. D. Santangelo, Phys. Rev.
Lett. 2012,109, 114301.
[93] E. D. Demaine, M. L. Demaine, V. Hart, G. N. Price, T. Tachi, in
Graphs and Combinatorics, Springer, Berlin 2011.
[94] T.-U. Lee, Y. Chen, M. T. Heitzmann, J. M. Gattas, Mater. Des. 2021,
207, 109859.
[95] N. Lee, S. Pellegrino, in Spacecraft Struct. Conf., AIAA, National Har-
bor, Maryland, USA, 2014.
[96] A. Körner, L. Born, A. Mader, R. Sachse, S. Saffarian, A. S.
Westermeier, S. Poppinga, M. Bischoff, G. T. Gresser, M. Milwich,
T. Speck, J. Knippers, Smart Mater. Struct. 2018,27, 017001.
[97] Y. Du, C. Song, J. Xiong, L. Wu, Compos. Sci. Technol. 2019,174,
94.
[98] Z. Zhai, Y. Wang, K. Lin, L. Wu, H. Jiang, Sci. Adv. 2020,6, eabe2000.
[99] Z.Yan,F.Zhang,J.Wang,F.Liu,X.Guo,K.Nan,Q.Lin,M.Gao,D.
Xiao, Y. Shi, Y. Qiu, H. Luan, J. H. Kim, Y. Wang, H. Luo, M. Han, Y.
Huang, Y. Zhang, J. A. Rogers, Adv. Funct. Mater. 2016,26, 2629.
[100] Z. Yan, F. Zhang, F. Liu, M. Han, D. Ou, Y. Liu, Q. Lin, X. Guo, H. Fu,
Z. Xie, M. Gao, Y. Huang, J. Kim, Y. Qiu, K. Nan, J. Kim, P. Gutruf,
H. Luo, An Zhao, K.-C. Hwang, Y. Huang, Y. Zhang, J. A. Rogers, Sci.
Adv. 2016,2, e1601014.
[101] T. Van Manen, S. Janbaz, M. Ganjian, A. A. Zadpoor, Mater. Today
2020,32, 59.
[102] M. Redoutey, A. Roy, E. T. Filipov, Int. J. Solids Struct. 2021,229,
111140.
[103] J. P. Whitney, P. S. Sreetharan, K. Y. Ma, R. J. Wood, J. Micromech.
Microeng. 2011,21, 115021.
Adv. Mater. 2023, 2302066 © 2023 Wiley-VCH GmbH
2302066 (28 of 35)
15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202302066 by Stanford University, Wiley Online Library on [12/12/2023]. 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 www.advmat.de
[104] Z. Liu, H. Du, J. Li, L. Lu, Z.-Y. Li, N. X. Fang, Sci. Adv. 2018,4,
eaat4436.
[105] H. Fu, K. Nan, W. Bai, W. Huang, K. Bai, L. Lu, C. Zhou, Y. Liu, F. Liu,
J. Wang, M. Han, Z. Yan, H. Luan, Y. Zhang, Y. Zhang, J. Zhao, Xu
Cheng, M. Li, J. W. Lee, Y. Liu, D. Fang, X. Li, Y. Huang, Y. Zhang, J.
A. Rogers, Nat. Mater. 2018,17, 268.
[106] F. Zhang, S. Li, Z. Shen, X. Cheng, Z. Xue, H. Zhang, H. Song, K.
Bai, D. Yan, H. Wang, Y. Zhang, Y. Huang, Proc. Natl. Acad. Sci. USA
2021,118, e2026414118.
[107] M. Han, X. Guo, X. Chen, C. Liang, H. Zhao, Q. Zhang, W. Bai, F.
Zhang,H.Wei,C.Wu,Q.Cui,S.Yao,B.Sun,Y.Yang,Q.Yang,Y.
Ma, Z. Xue, J. W. Kwak, T. Jin, Q. Tu, E. Song, Z. Tian, Y. Mei, D.
Fang, H. Zhang, Y. Huang, Y. Zhang, J. A. Rogers, Sci. Rob. 2022,7,
eabn0602.
[108] B. H. Kim, K. Li, J.-T. Kim, Y. Park, H. Jang, X. Wang, Z. Xie, S. M.
Won, H.-J. Yoon, G. Lee, W. J. Jang, K. H. Lee, T. S. Chung, Y. H.
Jung, S. Y. Heo, Y. Lee, J. Kim, T. Cai, Y. Kim, P. Prasopsukh, Y. Yu, X.
Yu, R. Avila, H. Luan, H. Song, F. Zhu, Y. Zhao, L. Chen, S. H. Han,
J. Kim, et al., Nature 2021,597, 503.
[109] K. Liu, F. Hacker, C. Daraio, Sci. Rob. 2021,6, eabf5116.
[110] X. Zang, C. Shen, Y. Chu, B. Li, M. Wei, J. Zhong, M. Sanghadasa, L.
Lin, Adv. Mater. 2018,30, 1800062.
[111] J. Ding, B. Li, L. Chen, W. Qin, Angew. Chem., Int. Ed. 2016,55, 13033.
[112] X. Chen, Y. Li, X. Wang, H. Yu, ACS Appl. Mater. Interfaces 2022,14,
36227.
[113] Y. Chen, R. Peng, Z. You, Science 2015,349, 396.
[114] T. Tachi, in Origami5 (Eds: P. Wang-Iverson, R. J. Lang, M. Yim),
Taylor & Francis Group, New York 2011, Ch. 20.
[115] M. Behl, M. Y. Razzaq, A. Lendlein, Adv. Mater. 2010,22, 3388.
[116] M. T. Tolley, S. M. Felton, S. Miyashita, D. Aukes, D. Rus, R. J. Wood,
Smart Mater. Struct. 2014,23, 094006.
[117] Y. Mao, K. Yu, M. S. Isakov, J. Wu, M. L. Dunn, H. J. Qi, Sci. Rep.
2015,5, 13616.
[118] S. Janbaz, R. Hedayati, A. A. Zadpoor, Mater. Horiz. 2016,3, 536.
[119] M. Wagner, T. Chen, K. Shea, 3D Print. Addit. Manuf. 2017,4, 133.
[120] A. Oyefusi, J. Chen, Angew. Chem., Int. Ed. 2017,56, 8250.
[121] J. Cui, J. G. M. Adams, Y. Zhu, Smart Mater. Struct. 2017,26, 125011.
[122] J. Cui, F. R. Poblete, Y. Zhu, Adv. Funct. Mater. 2018,28, 1802768.
[123] G. Li, S. Wang, Z. Liu, Z. Liu, H. Xia, C. Zhang, X. Lu, J. Jiang, Y.
Zhao, ACS Appl. Mater. Interfaces 2018,10, 40189.
[124] D. H. Wang, L.-S. Tan, ACS Macro Lett. 2019,8, 546.
[125] Y. Tang, Y. Li, Y. Hong, S. Yang, J. Yin, Proc. Natl. Acad. Sci. USA 2019,
116, 26407.
[126] X. Xin, L. Liu, Y. Liu, J. Leng, Smart Mater. Struct. 2020,29, 065015.
[127] R. Tao, L. Ji, Y. Li, Z. Wan, W. Hu, W. Wu, B. Liao, L. Ma, D. Fang,
Composites, Part B 2020,201, 108344.
[128] S. Jape, M. Garza, J. Ruff, F. Espinal, D. Sessions, G. Huff, D. C.
Lagoudas, E. A. Peraza Hernandez, D. J. Hartl, Smart Mater. Struct.
2020,29, 115011.
[129] C. Yang, M. Boorugu, A. Dopp, J. Ren, R. Martin, D. Han, W. Choi,
H. Lee, Mater. Horiz. 2019,6, 1244.
[130] L. Zhao, L. Wang, J. Shi, X. Hou, Qi Wang, Y. Zhang, Y. Wang, N. Bai,
J. Yang, J. Zhang, B. Yu, C. F. Guo, ACS Nano 2021,15, 5752.
[131] T. Langford, A. Mohammed, K. Essa, A. Elshaer, H. Hassanin, Appl.
Sci. 2021,11, 332.
[132] W. Zhao, N. Li, L. Liu, J. Leng, Y. Liu, Compos. Struct. 2022,293,
115669.
[133] S. Miyashita, L. Meeker, M. T. Tolley, R. J. Wood, D. Rus, Smart
Mater. Struct. 2014,23, 094005.
[134] J. Zhou, S. A. Turner, S. M. Brosnan, Q. Li, J.-M. Y. Carrillo, D.
Nykypanchuk, O. Gang, V. S. Ashby, A. V. Dobrynin, S. S. Sheiko,
Macromolecules 2014,47, 1768.
[135] C. Yuan, T. Wang, M. L. Dunn, H. J. Qi, Int. J. Precis. Eng. Manuf.-
Green Technol. 2017,4, 281.
[136] J.-E. Suh, Y. Miyazawa, J. Yang, J.-H. Han, Adv. Eng. Mater. 2022,24,
2101202.
[137] H. Yang, W. R. Leow, T. Wang, J. Wang, J. Yu, K. He, D. Qi, C. Wan,
X. Chen, Adv. Mater. 2017,29, 1701627.
[138] Q. Zhang, J. Wommer, C. O’rourke, J. Teitelman, Y. Tang, J. Robison,
G. Lin, J. Yin, Extreme Mech. Lett. 2017,11, 111.
[139] Y. Liu, J. K. Boyles, J. Genzer, M. D. Dickey, Soft Matter 2012,8, 1764.
[140] Y. Liu, B. Shaw, M. D. Dickey, J. Genzer, Sci. Adv. 2017,3, e1602417.
[141] J. R. Kumpfer, S. J. Rowan, J. Am. Chem. Soc. 2011,133, 12866.
[142] Z. Li, X. Zhang, S. Wang, Y. Yang, B. Qin, K. Wang, T. Xie, Y. Wei, Y.
Ji, Chem. Sci. 2016,7, 4741.
[143] Y. Lee, H. Lee, T. Hwang, J.-G. Lee, M. Cho, Sci. Rep. 2015,5, 16544.
[144] J. Xue, Y. Ge, Z. Liu, Z. Liu, J. Jiang, G. Li, ACS Appl. Mater. Interfaces
2022,14, 10836.
[145] Y. Ge, H. Wang, J. Xue, J. Jiang, Z. Liu, Z. Liu, G. Li, Y. Zhao, ACS
Appl. Mater. Interfaces 2021,13, 38773.
[146] M. Jamal, S. S. Kadam, R. Xiao, F. Jivan, T.-M. Onn, R. Fernandes, T.
D.Nguyen,D.H.Gracias,Adv. Healthcare Mater. 2013,2, 1142.
[147] J.-H. Na, A. A. Evans, J. Bae, M. C. Chiappelli, C. D. Santangelo, R.
J. Lang, T. C. Hull, R. C. Hayward, Adv. Mater. 2015,27, 79.
[148] X. Zhang, C. L. Pint, M. H. Lee, B. E. Schubert, A. Jamshidi, K. Takei,
H.Ko,A.Gillies,R.Bardhan,J.J.Urban,M.Wu,R.Fearing,A.Javey,
Nano Lett. 2011,11, 3239.
[149] G. Stoychev, S. Turcaud, J. W. C. Dunlop, L. Ionov, Adv. Funct. Mater.
2013,23, 2295.
[150] J. Kim, J. A. Hanna, R. C. Hayward, C. D. Santangelo, Soft Matter
2012,8, 2375.
[151] C. Yao, Z. Liu, C. Yang, W. Wang, X.-J. Ju, R. Xie, L.-Y. Chu, Adv. Funct.
Mater. 2015,25, 2980.
[152] C. Ma, W. Lu, X. Yang, J. He, X. Le, L. Wang, J. Zhang, M. J. Serpe, Y.
Huang, T. Chen, Adv. Funct. Mater. 2018,28, 1704568.
[153] E. Wang, M. S. Desai, S.-W. Lee, Nano Lett. 2013,13, 2826.
[154] Z. Lei, W. Zhu, S. Sun, P. Wu, Nanoscale 2016,8, 18800.
[155] D. Jin, Q. Chen, T.-Y. Huang, J. Huang, L. Zhang, H. Duan, Mater.
Today 2020,32, 19.
[156] N. Bassik, B. T. Abebe, K. E. Laflin, D. H. Gracias, Polymer 2010,51,
6093.
[157] L. Zhao, J. Huang, Y. Zhang, T. Wang, W. Sun, Z. Tong, ACS Appl.
Mater. Interfaces 2017,9, 11866.
[158] H. Deng, X. Xu, C. Zhang, J.-W. Su, G. Huang, J. Lin, ACS Appl. Mater.
Interfaces 2020,12, 13378.
[159] H. Zhang, X. Guo, J. Wu, D. Fang, Y. Zhang, Sci. Adv. 2018,4,
eaar8535.
[160] H. He, J. Guan, J. L. Lee, J. Controlled Release 2006,110, 339.
[161] J. C. Breger, C. K. Yoon, R. Xiao, H. R. Kwag, M. O. Wang, J. P. Fisher,
T. D. Nguyen, D. H. Gracias, ACS Appl. Mater. Interfaces 2015,7,
3398.
[162] K. Malachowski, J. Breger, H. R. Kwag, M. O. Wang, J. P. Fisher,
F. M. Selaru, D. H. Gracias, Angew. Chem., Int. Ed. 2014,53,
8045.
[163] T. S. Shim, S.-H. Kim, C.-J. Heo, H. C. Jeon, S.-M. Yang, Angew.
Chem., Int. Ed. 2012,51, 1420.
[164] G. Stoychev, N. Puretskiy, L. Ionov, Soft Matter 2011,7, 3277.
[165] Y. Li, Y. Teixeira, G. Parlato, J. Grace, F. Wang, B. D. Huey, X. Wang,
Soft Matter 2022, 6857.
[166] G. Chen, B. Jin, Y. Shi, Q. Zhao, Y. Shen, T. Xie, Adv. Mater. 34,
2201679.
[167] L. Ren, B. Li, Y. He, Z. Song, X. Zhou, Q. Liu, Q. Liu, L. Ren, ACS
Appl. Mater. Interfaces 2020,12, 15562.
[168] Y. Huang, H. K. Bisoyi, S. Huang, M. Wang, X.-M. Chen, Z. Liu, H.
Yang, Q. Li , Angew. Chem., Int. Ed. 2021,60, 11247.
[169] M. K. Mcbride, A. M. Martinez, L. Cox, M. Alim, K. Childress, M.
Beiswinger, M. Podgorski, B. T. Worrell, J. Killgore, C. N. Bowman,
Sci. Adv. 2018,4, eaat4634.
Adv. Mater. 2023, 2302066 © 2023 Wiley-VCH GmbH
2302066 (29 of 35)
15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202302066 by Stanford University, Wiley Online Library on [12/12/2023]. 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 www.advmat.de
[170] Z.-C. Jiang, Y.-Y. Xiao, X. Tong, Y. Zhao, Angew. Chem., Int. Ed. 2019,
58, 5332.
[171] K. Fuchi, T. H. Ware, P. R. Buskohl, G. W. Reich, R. A. Vaia, T. J. White,
J. J. Joo, Soft Matter 2015,11, 7288.
[172] A. Kotikian, C. Mcmahan, E. C. Davidson, J. M. Muhammad, R. D.
Weeks, C. Daraio, J. A. Lewis, Sci. Rob. 2019,4, eaax7044.
[173] T. H. Ware, M. E. Mcconney, J. J. Wie, V. P. Tondiglia, T. J. White,
Science 2015,347, 982.
[174] L. T. De Haan, J. M. N. Verjans, D. J. Broer, C. W. M. Bastiaansen, A.
P. H. J. Schenning, J. Am. Chem. Soc. 2014,136, 10585.
[175] K. Kim, Y. Guo, J. Bae, S. Choi, H. Y. Song, S. Park, K. Hyun, S.-K.
Ahn, Small 2021,17, 2100910.
[176] H. Kim, J. A. Lee, C. P. Ambulo, H. B. Lee, S. H. Kim, V. V. Naik, C. S.
Haines, A. E. Aliev, R. Ovalle-Robles, R. H. Baughman, T. H. Ware,
Adv. Funct. Mater. 2019,29, 1905063.
[177] R. C. P. Verpaalen, M. Pilz Da Cunha, T. A. P. Engels, M. G. Debije,
A. P. H. J. Schenning, Angew. Chem., Int. Ed. 2020,59, 4532.
[178] X. Pang, L. Qin, B. Xu, Q. Liu, Y. Yu, Adv. Funct. Mater. 2020,30,
2002451.
[179] X. Peng, S. Wu, X. Sun, L. Yue, S. M. Montgomery, F. Demoly, K.
Zhou, R. R. Zhao, H. J. Qi, Adv. Mater. 2022,34, 2204890.
[180] D. J. Roach, X. Kuang, C. Yuan, K. Chen, H. J. Qi, Smart Mater. Struct.
2018,27, 125011.
[181] V. Maurin, Y. Chang, Q. Ze, S. Leanza, R. R. Zhao, (Preprint)
arXiv:2302. 13583, submitted: Feb 2023.
[182] S. Yi, L. Wang, Z. Chen, J. Wang, X. Song, P. Liu, Y. Zhang, Q. Luo,
L.Peng,Z.Wu,C.F.Guo,L.Jiang,Nat. Commun. 2022,13, 4177.
[183] Z. Ren, W. Hu, X. Dong, M. Sitti, Nat. Commun. 2019,10, 2703.
[184] J. Cui, T.-Y. Huang, Z. Luo, P. Testa, H. Gu, X.-Z. Chen, B. J. Nelson,
L. J. Heyderman, Nature 2019,575, 164.
[185] S. Wu, Q. Ze, J. Dai, N. Udipi, G. H. Paulino, R. Zhao, Proc. Natl.
Acad. Sci. USA 2021,118, e2110023118.
[186] T. G. Leong, C. L. Randall, B. R. Benson, N. Bassik, G. M. Stern, D.
H. Gracias, Proc. Natl. Acad. Sci. USA 2009,106, 703.
[187] L. Li, H. Yao, S. Mi, ACS Appl. Mater. Interfaces 2023,15, 3486.
[188] D. Tang, C. Zhang, H. Sun, H. Dai, J. Xie, J. Fu, P. Zhao, Nano Energy
2021,89, 106424.
[189] Y. Alapan, A. C. Karacakol, S. N. Guzelhan, I. Isik, M. Sitti, Sci. Adv.
2020,6, eabc6414.
[190] H. Deng, K. Sattari, Y. Xie, P. Liao, Z. Yan, J. Lin, Nat. Commun. 2020,
11, 6325.
[191] B. S. Yeow, H. Yang, M. S. Kalairaj, H. Gao, C. J. Cai, S. Xu, P.-Y. Chen,
H. Ren, Adv. Mater. Technol. 2022,7, 2101140.
[192] T. Xu, J. Zhang, M. Salehizadeh, O. Onaizah, E. Diller, Sci. Rob. 2019,
4, eaav4494.
[193] Y. Liu, H. Du, L. Liu, J. Leng, Smart Mater. Struct. 2014,23, 023001.
[194] W. Zhao, L. Liu, F. Zhang, J. Leng, Y. Liu, Mater. Sci. Eng., C 2019,97,
864.
[195] H. Gao, J. Li, F. Zhang, Y. Liu, J. Leng, Mater. Horiz. 2019,6, 931.
[196] Q. Zhao, H. J. Qi, T. Xie, Prog. Polym. Sci. 2015,49, 79.
[197] K. Yu, Q. Ge, H. J. Qi, Nat. Commun. 2014,5, 3066.
[198] A. Lendlein, S. Kelch, Angew. Chem., Int. Ed. 2002,41, 2034.
[199] M. Behl, K. Kratz, J. Zotzmann, U. Nöchel, A. Lendlein, Adv. Mater.
2013,25, 4466.
[200] T. Chung, A. Romo-Uribe, P. T. Mather, Macromolecules 2008,41,
184.
[201] K. K. Westbrook, P. T. Mather, V. Parakh, M. L. Dunn, Q. Ge, B. M.
Lee, H. J. Qi, Smart Mater. Struct. 2011,20, 065010.
[202] I. Bellin, S. Kelch, R. Langer, A. Lendlein, Proc. Natl. Acad. Sci. USA
2006,103, 18043.
[203] T. Xie, Nature 2010,464, 267.
[204] Y. Mao, Z. Ding, C. Yuan, S. Ai, M. Isakov, J. Wu, T. Wang, M. L.
Dunn, H. J. Qi, Sci. Rep. 2016,6, 24761.
[205] C. Yuan, F. Wang, Q. Ge, Extreme Mech. Lett. 2021,42, 101122.
[206] J. A.-C. Liu, J. H. Gillen, S. R. Mishra, B. A. Evans, J. B. Tracy, Sci. Adv.
2019,5, eaaw2897.
[207] Q. Ze, X. Kuang, S. Wu, J. Wong, S. M. Montgomery, R. Zhang, J. M.
Kovitz, F. Yang, H. J. Qi, R. Zhao, Adv. Mater. 32, 1906657.
[208] D. J. Roach, X. Sun, X. Peng, F. Demoly, K. Zhou, H. J. Qi, Adv. Funct.
Mater. 2022,32, 2203236.
[209] S. Wu, J. Eichenberger, J. Dai, Y. Chang, N. Ghalichechian, R. R.
Zhao, Adv. Intell. Syst. 2022,4, 2200106.
[210] L. Cera, G. M. Gonzalez, Q. Liu, S. Choi, C. O. Chantre, J. Lee, R.
Gabardi, M. C. Choi, K. Shin, K. K. Parker, Nat. Mater. 20, 242.
[211] A. Firouzeh, M. Salerno, J. Paik, IEEE Trans. Robot. 2017,33, 765.
[212] A. Firouzeh, J. Paik, IEEE ASME Trans. Mechatron. 2017,22, 2165.
[213] X. Liu, J. Liu, S. Lin, X. Zhao, Mater. Today 2020,36, 102.
[214] P. Calvert, Adv. Mater. 2009,21, 743.
[215] Y. Ma, M. Hua, S. Wu, Y. Du, X. Pei, X. Zhu, F. Zhou, X. He, Sci. Adv.
2020,6, eabd2520.
[216] G. Stoychev, L. Guiducci, S. Turcaud, J. W. C. Dunlop, L. Ionov, Adv.
Funct. Mater. 2016,26, 7733.
[217] H. Na, Y.-W. Kang, C. S. Park, S. Jung, H.-Y. Kim, J.-Y. Sun, Science
2022,376, 301.
[218] V. R. Feig, H. Tran, M. Lee, Z. Bao, Nat. Commun. 2018,9, 2740.
[219] S. J. Wu, H. Yuk, J. Wu, C. S. Nabzdyk, X. Zhao, Adv. Mater. 33,
2007667.
[220] Y. Pan, Z. Yang, C. Li, S. U. Hassan, H. C. Shum, Sci. Adv. 2022,8,
eabo1719.
[221] D. Morales, E. Palleau, M. D. Dickey, O. D. Velev, Soft Matter 2014,
10, 1337.
[222] C. Yang, W. Wang, C. Yao, R. Xie, X.-J. Ju, Z. Liu, L.-Y. Chu, Sci. Rep.
2015,5, 13622.
[223] R. Luo, J. Wu, N.-D. Dinh, C.-H. Chen, Adv. Funct. Mater. 2015,25,
7272.
[224] K. Otake, H. Inomata, M. Konno, S. Saito, Macromolecules 1990,23,
283.
[225] T. Tanaka, Phys. Rev. Lett. 1978,40, 820.
[226] J. Kim, J. A. Hanna, M. Byun, C. D. Santangelo, R. C. Hayward, Sci-
ence 2012,335, 1201.
[227] G. Stoychev, S. Zakharchenko, S. Turcaud, J. W. C. Dunlop, L. Ionov,
ACS Nano 2012,6, 3925.
[228] J. Wang, J. Wang, Z. Chen, S. Fang, Y. Zhu, R. H. Baughman, L. Jiang,
Chem. Mater. 2017,29, 9793.
[229] X. Peng, C. Jiao, Y. Zhao, N. Chen, Y. Wu, T. Liu, H. Wang, ACS Appl.
Nano Mater. 2018,1, 1522.
[230] Q. Zhao, Y. Liang, L. Ren, Z. Yu, Z. Zhang, L. Ren, Nano Energy 2018,
51, 621.
[231] M. E. Lee-Trimble, J.-H. Kang, R. C. Hayward, C. D. Santangelo, Soft
Matter 2022,18, 6384.
[232] J.-H Kang, H. Kim, C. D. Santangelo, R. C. Hayward, Adv. Mater.
2019,31, 0193006.
[233] M. Trujillo-Miranda, I. Apsite, J. A. R. Agudo, G. Constante, L. Ionov,
Macromol. Biosci. 2023,23, 2200320.
[234] G. Go, V. D. Nguyen, Z. Jin, J.-O. Park, S. Park, Int. J. Control Autom.
Syst. 2018,16, 1341.
[235] Z. Li, P. Liu, X. Ji, J. Gong, Y. Hu, W. Wu, X. Wang, H.-Q. Peng, R. T.
K.Kwok,J.W.Y.Lam,J.Lu,B.Z.Tang,Adv. Mater. 32, 1906493.
[236] A. Aggarwal, C. Li, S. I. Stupp, M. Olvera De La Cruz, Soft Matter
2022,18, 2193.
[237] C. Li, Y. Xue, M. Han, L. C. Palmer, J. A. Rogers, Y. Huang, S. I. Stupp,
Matter 2021,4, 1377.
[238] Z. Chen, Y. Li, Q. M. Li, Mater. Des. 2021,207, 109819.
[239] H. Wermter, H. Finkelmann, e-Polymers 2001,1, 013.
[240] J. Küpfer, H. Finkelmann, Makromol. Chem., Rapid Commun. 1991,
12, 717.
[241] D. L. Thomsen, P. Keller, J. Naciri, R. Pink, H. Jeon, D. Shenoy, B. R.
Ratna, Macromolecules 2001,34, 5868.
Adv. Mater. 2023, 2302066 © 2023 Wiley-VCH GmbH
2302066 (30 of 35)
15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202302066 by Stanford University, Wiley Online Library on [12/12/2023]. 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 www.advmat.de
[242] Y. Yu, M. Nakano, T. Ikeda, Nature 2003,425, 145.
[243] M. Camacho-Lopez, H. Finkelmann, P. Palffy-Muhoray, M. Shelley,
Nat. Mater. 2004,3, 307.
[244] K. D. Harris, C. W. M. Bastiaansen, J. Lub, D. J. Broer, Nano Lett.
2005,5, 1857.
[245] M. Brehmer, R. Zentel, G. Wagenblast, K. Siemensmeyer, Macromol.
Chem. Phys. 1994,195, 1891.
[246] C. M. Spillmann, B. R. Ratna, J. Naciri, Appl. Phys. Lett. 2007,90,
021911.
[247] A. Kaiser, M. Winkler, S. Krause, H. Finkelmann, A. M. Schmidt, J.
Mater. Chem. 2009,19, 538.
[248] P. G. de Gennes, J. Prost, in The Physics of Liquid Crystals, Clarendon
Press, Oxford 1993.
[249] K. M. Herbert, H. E. Fowler, J. M. Mccracken, K. R. Schlafmann, J.
A. Koch, T. J. White, Nat. Rev. Mater. 2022,7, 23.
[250] M. Behl, A. Lendlein, Soft Matter 2007,3, 58.
[251] H. Finkelmann, H.-J. Kock, G. Rehage, Makromol. Chem., Rapid
Commun. 1981,2, 317.
[252] M. Warner, E. M. Terentjev, in Liquid Crystal Elastomers, Vol. 120,
Oxford University Press, Oxford 2003.
[253] S. W. Ula, N. A. Traugutt, R. H. Volpe, R. R. Patel, K. Yu, C. M. Yakacki,
Liq. Cryst. Rev. 2018,6, 78.
[254] P.-G. De Gennes, M. Hébert, R. Kant, Macromol. Symp. 1997,113,
39.
[255] D. J. Roach, C. Yuan, X. Kuang, V. C.-F. Li, P. Blake, M. L. Romero, I.
Hammel, K. Yu, H. J. Qi, ACS Appl. Mater. Interfaces 2019,11, 19514.
[256] C. L. Van Oosten, K. D. Harris, C. W. M. Bastiaansen, D. J. Broer,
Eur. Phys. J. Spec. Top. 2007,23, 329.
[257] X. Lu, S. Guo, X. Tong, H. Xia, Y. Zhao, Adv. Mater. 2017,29, 1606467.
[258] R. A. M. Hikmet, D. J. Broer, Polymer 1991,32, 1627.
[259] C. Ohm, M. Brehmer, R. Zentel, Adv. Mater. 2010,22, 3366.
[260] K.-W. Lee, S.-H. Paek, A. Lien, C. Durning, H. Fukuro, Macro-
molecules 1996,29, 8894.
[261] S. Krause, R. Dersch, J. H. Wendorff, H. Finkelmann, Macromol.
Rapid Commun. 2007,28, 2062.
[262] C. M. Yakacki, M. Saed, D. P. Nair, T. Gong, S. M. Reed, C. N.
Bowman, RSC Adv. 2015,5, 18997.
[263] K. Mehta, A. R. Peeketi, L. Liu, D. Broer, P. Onck, R. K. Annabattula,
Appl. Phys. Rev. 2020,7, 041306.
[264] G. N. Mol, K. D. Harris, C. W. M. Bastiaansen, D. J. Broer, Adv. Funct.
Mater. 2005,15, 1155.
[265] J. M. Boothby, T. H. Ware, Soft Matter 2017,13, 4349.
[266] M. Warner, C. D. Modes, D. Corbett, Proc. R. Soc. A 2010,466, 2975.
[267] L. T. De Haan, A. P. H. J. Schenning, D. J. Broer, Polymer 2014,55,
5885.
[268] B. R. Donovan, V. M. Matavulj, S.-K. Ahn, T. Guin, T. J. White, Adv.
Mater. 2019,31, 1805750.
[269] Yu Xia, G. Cedillo-Servin, R. D. Kamien, S. Yang, Adv. Mater. 2016,
28, 9637.
[270] S.-K. Ahn, T. H. Ware, K. M. Lee, V. P. Tondiglia, T. J. White, Adv. Funct.
Mater. 2016,26, 5819.
[271] H. Aharoni, Y. Xia, X. Zhang, R. D. Kamien, S. Yang, Proc. Natl. Acad.
Sci. USA 2018,115, 7206.
[272] Z. S. Davidson, H. Shahsavan, A. Aghakhani, Y. Guo, L. Hines, Y.
Xia, S. Yang, M. Sitti, Sci. Adv. 2019,5, eaay0855.
[273] A. F. Minori, Q. He, P. E. Glick, I. Adibnazari, A. Stopol, S. Cai, M. T.
Tolley, Smart Mater. Struct. 2020,29, 105003.
[274] M. Zhang, H. Shahsavan, Y. Guo, A. Pena-Francesch, Y. Zhang, M.
Sitti, Adv. Mater. 2021,33, 2008605.
[275] J. Wu, S. Yao, H. Zhang, W. Man, Z. Bai, F. Zhang, X. Wang, D. Fang,
Y. Z h a n g , Adv. Mater. 2021,33, 2106175.
[276] W. Hu, G. Z. Lum, M. Mastrangeli, M. Sitti, Nature 2018,554, 81.
[277] W. Xi, A. A. Solovev, A. N. Ananth, D. H. Gracias, S. Sanchez, O. G.
Schmidt, Nanoscale 2013,5, 1294.
[278] R. Zhang, S. Wu, Q. Ze, R. Zhao, J. Appl. Mech. 2020,87, 091008.
[279] H. Song, H. Lee, J. Lee, J. K. Choe, S. Lee, J. Y. Yi, S. Park, J.-W. Yoo,
M. S. Kwon, J. Kim, Nano Lett. 2020,20, 5185.
[280] A. Ghosh, L. Li, L. Xu, R. P. Dash, N. Gupta, J. Lam, Q. Jin, V.
Akshintala, G. Pahapale, W. Liu, A. Sarkar, R. Rais, D. H. Gracias,
F. M. Selaru , Sci. Adv. 2020,6, eabb4133.
[281] C. J. Kloxin, T. F. Scott, B. J. Adzima, C. N. Bowman, Macromolecules
2010,43, 2643.
[282] N. Zheng, Y. Xu, Q. Zhao, T. Xie, Chem. Rev. 2021,121, 1716.
[283] D. Montarnal, M. Capelot, F. Tournilhac, L. Leibler, Science 2011,
334, 965.
[284] P. Zheng, T. J. Mccarthy, J. Am. Chem. Soc. 2012,134, 2024.
[285] X. Chen, M. A. Dam, K. Ono, A. Mal, H. Shen, S. R. Nutt, K. Sheran,
F. Wu d l , Science 2002,295, 1698.
[286] T. F. Scott, A. D. Schneider, W. D. Cook, C. N. Bowman, Science 2005,
308, 1615.
[287] A. M. Peterson, R. E. Jensen, G. R. Palmese, ACS Appl. Mater. Inter-
faces 2010,2, 1141.
[288] X. He, Y. Lin, Y. Ding, A. M. Abdullah, Z. Lei, Y. Han, X. Shi, W. Zhang,
K. Yu, Int. J. Extreme Manuf. 2021,4, 015301.
[289] J. Deng, X. Kuang, R. Liu, W. Ding, A. C. Wang, Y.-C. Lai, K. Dong, Z.
Wen, Y. Wang, L. Wang, H. J. Qi, T. Zhang, Z. L. Wang, Adv. Mater.
2018,30, 1705918.
[290] W. Denissen, J. M. Winne, F. E. Du Prez, Chem. Sci. 2016,7, 30.
[291] B. R. Elling, W. R. Dichtel, ACS Cent. Sci. 2020,6, 1488.
[292] Q. Zhao, W. Zou, Y. Luo, T. Xie, Sci. Adv. 2016,2, e1501297.
[293] G. Zhang, Q. Zhao, L. Yang, W. Zou, X. Xi, T. Xie, ACS Macro Lett.
2016,5, 805.
[294] B. Jin, H. Song, R. Jiang, J. Song, Q. Zhao, T. Xie, Sci. Adv. 2018,4,
eaao3865.
[295] Z. Ding, Li Yuan, G. Liang, A. Gu, J. Mater. Chem. A 2019,7, 9736.
[296] J. Zhou, H. Yue, M. Huang, C. Hao, S. He, H. Liu, W. Liu, C. Zhu, X.
Dong, D. Wang, ACS Appl. Mater. Interfaces 2021,13, 43426.
[297] Z. Pei, Y. Yang, Q. Chen, E. M. Terentjev, Y. Wei, Y. Ji, Nat. Mater.
2014,13, 36.
[298] Z. Pei, Y. Yang, Q. Chen, Y. Wei, Y. Ji, Adv. Mater. 2016,28,156.
[299] Y. Yang, Z. Pei, Z. Li, Y. Wei, Y. Ji, J. Am. Chem. Soc. 2016,138, 2118.
[300] Z. Li, Y. Yang, Z. Wang, X. Zhang, Q. Chen, X. Qian, N. Liu, Y. Wei,
Y. J i , J. Mater. Chem. A 2017,5, 6740.
[301] Y. Yang, E. M. Terentjev, Y. Wei, Y. Ji, Nat. Commun. 2018,9, 1906.
[302] Y. Wu, Y. Yang, X. Qian, Q. Chen, Y. Wei, Y. Ji, Angew. Chem., Int. Ed.
2020,59, 4778.
[303] M. K. McBride, M. Podgorski, S. Chatani, B. T. Worrell, C. N.
Bowman, ACS Appl. Mater. Interfaces 2018,10, 22739.
[304] Y. Li, O. Rios, J. K. Keum, J. Chen, M. R. Kessler, ACS Appl. Mater.
Interfaces 2016,8, 15750.
[305] Y. Li, Y. Zhang, O. Rios, J. K. Keum, M. R. Kessler, RSC Adv. 2017,7,
37248.
[306] X. Lu, H. Zhang, G. Fei, B. Yu, X. Tong, H. Xia, Y. Zhao, Adv. Mater.
2018,30, 1706597.
[307] X. Kuang, S. Wu, Q. Ze, L. Yue, Yi Jin, S. M. Montgomery, F. Yang,
H. J. Qi, R. Zhao, Adv. Mater. 2021,33, 2102113.
[308] H. Qiu, S. Wei, H. Liu, B. Zhan, H. Yan, W. Lu, J. Zhang, S. Wu, T.
Chen, Adv. Intell. Syst. 2021,3, 2000239.
[309] Y. Zhang, X. Le, Y. Jian, W. Lu, J. Zhang, T. Chen, Adv. Funct. Mater.
2019,29, 1905514.
[310] H. Okuzaki, T. Saido, H. Suzuki, Y. Hara, H. Yan, J. Phys.: Conf. Ser.
2007,127, 012001.
[311] D.-D. Han, Y.-L. Zhang, H.-B. Jiang, H. Xia, J. Feng, Q.-D. Chen, H.-L.
Xu, H.-B. Sun, Adv. Mater. 2015,27, 332.
[312] J. Mu, C. Hou, H. Wang, Y. Li, Q. Zhang, M. Zhu, Sci. Adv. 2015,1,
e1500533.
[313] W. Xu, Z. Qin, C.-T.Chen, H. R. Kwag, Q. Ma, A. Sarkar, M. J. Buehler,
D. H. Gracias, Sci. Adv. 2017,3, e1701084.
Adv. Mater. 2023, 2302066 © 2023 Wiley-VCH GmbH
2302066 (31 of 35)
15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202302066 by Stanford University, Wiley Online Library on [12/12/2023]. 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 www.advmat.de
[314] G. Cai, J.-H. Ciou, Y. Liu, Y. Jiang, P. S. Lee, Sci. Adv. 2019,5,
eaaw7956.
[315] Q. Zhao, J. W. C. Dunlop, X. Qiu, F. Huang, Z. Zhang, J. Heyda, J.
Dzubiella, M. Antonietti, J. Yuan, Nat. Commun. 2014,5, 4293.
[316] Xi Fan, W. Nie, H. Tsai, N. Wang, H. Huang, Y. Cheng, R. Wen, L.
Ma, F. Yan, Y. Xia, Adv. Sci. 2019,6, 1900813.
[317] S. Taccola, F. Greco, E. Sinibaldi, A. Mondini, B. Mazzolai, V. Mattoli,
Adv. Mater. 2015,27, 1668.
[318] K. E. Laflin, C. J. Morris, T. Muqeem, D. H. Gracias, Appl. Phys. Lett.
2012,101, 131901.
[319] D. Davis, R. Mailen, E. Luong, A. Russell, M. D. Dickey, J. Genzer,
ACS Appl. Eng. Mater. 2023,1, 193.
[320] D. Joung, A. Nemilentsau, K. Agarwal, C. Dai, C. Liu, Q. Su, J. Li, T.
Low, S. J. Koester, J.-H. Cho, Nano Lett. 2017,17, 1987.
[321] B. Han, Y.-L. Zhang, L. Zhu, Y. Li, Z.-C. Ma, Y.-Q. Liu, X.-L. Zhang,
X.-W. Cao, Q.-D. Chen, C.-W. Qiu, H.-B. Sun, Adv. Mater. 2019,31,
1970029.
[322] Z. Tang, Z. Gao, S. Jia, F. Wang, Y. Wang, Adv. Sci. 2017,4, 1600437.
[323] E. Hawkes, B. An, N. M. Benbernou, H. Tanaka, S. Kim, E. D.
Demaine, D. Rus, R. J. Wood, Proc. Natl. Acad. Sci. USA 2010,107,
12441.
[324] M. Boyvat, J.-S. Koh, R. J. Wood, Sci. Rob. 2017,2, eaan1544.
[325] P. K. Kumar, D. C. Lagoudas, in Shape Memory Alloys: Modeling and
Engineering Applications, Springer, Boston, MA 2008 p. 1.
[326] A. Rao, A. Srinivasa, J. Reddy, Design of Shape Memory Alloy (SMA)
Actuators,SpringerCham,2015.
[327] P. Velvaluri, A. Soor, P. Plucinsky, R. L. De Miranda, R. D. James, E.
Quandt, Sci. Rep. 2021,11, 10988.
[328] L. M. Fonseca, G. V. Rodrigues, M. A. Savi, A. Paiva, Chaos, Solitons
& Fractals 2019,122, 245.
[329] H. Yi, D. Kim, Y. Kim, D. Kim, J.-S. Koh, M.-J. Kim, Autom. Constr.
2020,114, 103151.
[330] C. D. Onal, R. J. Wood, D. Rus, in 2011 IEEE Int. Conf. on Robot. and
Autom., Shanghai, China 2011.
[331] N. Lazarus, G. L. Smith, M. D. Dickey, Adv. Intell. Syst. 2019,1,
1900059.
[332] A. L. Bachmann, B. Hanrahan, M. D. Dickey, N. Lazarus, ACS Appl.
Mater. Interfaces 2022,14, 14774.
[333] Y. Shi, F. Zhang, K. Nan, X. Wang, J. Wang, Y. Zhang, Y. Zhang, H.
Luan, K.-C. Hwang, Y. Huang, J. A. Rogers, Y. Zhang, Extreme Mech.
Lett. 2017,11, 105.
[334] Q. Liu, W. Wang, M. F. Reynolds, M. C. Cao, M. Z. Miskin, T. A. Arias,
D. A. Muller, P. L. Mceuen, I. Cohen, Sci. Rob. 2021,6, eabe6663.
[335] J. S. Randhawa, M. D. Keung, P. Tyagi, D. H. Gracias, Adv. Mater.
2010,22, 407.
[336] M. Taghavi, T. Helps, J. Rossiter, Sci. Rob. 2018,3, eaau9795.
[337] E. Acome, C. Keplinger, M. D. Gross, C. Bruns, D. Leithinger, in
Extended Abstracts of the 2021 CHI Conference on Human Factors
in Computing Systems, Association for Computing Machinery, Pur-
nendu, Yokohama, Japan 2021, Article 377.
[338] R. Pelrine, R. Kornbluh, Q. Pei, J. Joseph, Science 2000,287, 836.
[339] S. Shian, K. Bertoldi, D. R. Clarke, Adv. Mater. 2015,27, 6814.
[340] G. Kofod, W. Wirges, M. Paajanen, S. Bauer, Appl. Phys. Lett. 2007,
90, 081916.
[341] J. Shintake, S. Rosset, B. Schubert, D. Floreano, H. Shea, Adv. Mater.
2016,28, 231.
[342] M. Duduta, R. J. Wood, D. R. Clarke, Adv. Mater. 2016,28, 8058.
[343] J. Wang, S. Li, D. Gao, J. Xiong, P. S. Lee, NPG Asia Mater. 2019,11,
71.
[344] W. Pang, X. Cheng, H. Zhao, X. Guo, Z. Ji, G. Li, Y. Liang, Z. Xue, H.
Song, F. Zhang, Z. Xu, L. Sang, W. Huang, T. Li, Y. Zhang, Natl. Sci.
Rev. 2019,7, 342.
[345] J. Li, H. Godaba, Z. Q. Zhang, C. C. Foo, J. Zhu, Extreme Mech. Lett.
2018,24, 30.
[346] Y. Sun, D. Li, M. Wu, Y. Yang, J. Su, T. Wong, K. Xu, Y. Li, L. Li, X. Yu,
J. Yu, Microsyst. Nanoeng. 2022,8, 37.
[347] G. Liu, Y. Zhao, G. Wu, J. Lu, Sci. Adv. 2018,4, eaat0641.
[348] K. Huang, H. Elsayed, G. Franchin, P. Colombo, Addit. Manuf. 2020,
33, 101144.
[349] H. Jaffe, J. Am. Ceram. Soc. 1958,41, 494.
[350] K. Y. Ma, P. Chirarattananon, S. B. Fuller, R. J. Wood, Science 2013,
340, 603.
[351] Q. Zhao, S. Liu, J. Chen, G. He, J. Di, L. Zhao, T. Su, M. Zhang, Z.
Hou, Robot. Auton. Syst. 2021,140, 103733.
[352] J. S. Pulskamp, R. G. Polcawich, R. Q. Rudy, S. S. Bedair, R. M. Proie,
T. Ivanov, G. L. Smith, MRS Bull. 2012,37, 1062.
[353] H. Mcclintock, F. Z. Temel, N. Doshi, J.-S. Koh, R. J. Wood, Sci. Rob.
2018,3, eaar3018.
[354] S. Naficy, R. Gately, R. Gorkin, H. Xin, G. M. Spinks, Macromol.
Mater. Eng. 2017,302, 1600212.
[355] J. Guan, H. He, D. J. Hansford, L. J. Lee, J. Phys. Chem. B 2005,109,
23134.
[356] G. Z. Lum, Z. Ye, X. Dong, H. Marvi, O. Erin, W. Hu, M. Sitti, Proc.
Natl. Acad. Sci. USA 2016,113, E6007.
[357] S. Wu, Q. Ze, R. Zhang, N. Hu, Y. Cheng, F. Yang, R. Zhao, ACS Appl.
Mater. Interfaces 2019,11, 41649.
[358] D. Raviv, W. Zhao, C. Mcknelly, A. Papadopoulou, A. Kadambi, B.
Shi, S. Hirsch, D. Dikovsky, M. Zyracki, C. Olguin, R. Raskar, S.
Tibbits, Sci. Rep. 2014,4, 7422.
[359] Q. Ge, H. J. Qi, M. L. Dunn, Appl. Phys. Lett. 2013,103, 131901.
[360] J. A. Lewis, G. M. Gratson, Mater. Today 2004,7, 32.
[361] J. E. Smay, J. Cesarano, J. A. Lewis, Langmuir 2002,18, 5429.
[362] A. Sydney Gladman, E. A. Matsumoto, R. G. Nuzzo, L. Mahadevan,
J. A. Lewis, Nat. Mater. 2016,15, 413.
[363] J. Tang, B. Sun, Q. Yin, M. Yang, J. Hu, T. Wang, J. Mater. Chem. B
2021,9, 9183.
[364] P. Zhu, W. Yang, R. Wang, S. Gao, B. Li, Q. Li, ACS Appl. Mater. In-
terfaces 2018,10, 36435.
[365] Y. Zhang, Q. Wang, S. Yi, Z. Lin, C. Wang, Z. Chen, L. Jiang, ACS
Appl. Mater. Interfaces 2021,13, 4174.
[366] C. Ma, S. Wu, Q. Ze, X. Kuang, R. Zhang, H. J. Qi, R. Zhao, ACS Appl.
Mater. Interfaces 2021,13, 12639.
[367] S. Wu, C. M. Hamel, Q. Ze, F. Yang, H. J. Qi, R. Zhao, Adv. Intell.
Syst. 2020,2, 2000060.
[368] J. N. Rodriguez, C. Zhu, E. B. Duoss, T. S. Wilson, C. M. Spadaccini,
J. P. Lewicki, Sci. Rep. 2016,6, 27933.
[369] C. P. Ambulo, J. J. Burroughs, J. M. Boothby, H. Kim, M. R. Shankar,
T. H. Wa r e , ACS Appl. Mater. Interfaces 2017,9, 37332.
[370] A. Kotikian, R. L. Truby, J. W. Boley, T. J. White, J. A. Lewis, Adv. Mater.
2018,30, 1706164.
[371] M. Pozo, J. A. H. P. Sol, S. H. P. van Uden, A. R. Peeketi, S. J. D.
Lugger, R. K. Annabattula, A. P. H. J. Schenning, M. G. Debije, ACS
Appl. Mater. Interfaces 2021,13, 59381.
[372] C. Zhang, X. Lu, G. Fei, Z. Wang, H. Xia, Y. Zhao, ACS Appl. Mater.
Interfaces 2019,11, 44774.
[373] Z. Wang, Z. Wang, Y. Zheng, Q. He, Y. Wang, S. Cai, Sci. Adv. 2020,
6, eabc0034.
[374] X. Peng, X. Kuang, D. J. Roach, Y. Wang, C. M. Hamel, C. Lu, H. J.
Qi, Addit. Manuf. 2021,40, 101911.
[375] Y. Yang, Y. Chen, Y. Wei, Y. Li, Int. J. Adv. Manuf. Technol. 2016,84,
2079.
[376] S. Chen, Q. Zhang, J. Feng, J. Mater. Chem. C 2017,5, 8361.
[377] B. Peng, Y. Yang, T. Ju, K. A. Cavicchi, ACS Appl. Mater. Interfaces
2021,13, 12777.
[378] T. Van Manen, S. Janbaz, A. A. Zadpoor, Mater. Horiz. 2017,4, 1064.
[379] T. Van Manen, S. Janbaz, K. M. B. Jansen, A. A. Zadpoor, Commun.
Mater. 2021,2, 56.
[380] B. Zou, C. Song, Z. He, J. Ju, Extreme Mech. Lett. 2022,54, 101779.
Adv. Mater. 2023, 2302066 © 2023 Wiley-VCH GmbH
2302066 (32 of 35)
15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202302066 by Stanford University, Wiley Online Library on [12/12/2023]. 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 www.advmat.de
[381] A. B. Baker, S. R. G. Bates, T. M. Llewellyn-Jones, L. P. B. Valori, M.
P. M. Dicker, R . S . Tr a s k , Mater. Des. 2019,163, 107544.
[382] S. Qi, H. Guo, J. Fu, Y. Xie, M. Zhu, M. Yu, Compos. Sci. Technol.
2020,188, 107973.
[383] X. Cao, S. Xuan, S. Sun, Z. Xu, J. Li, X. Gong, ACS Appl. Mater. Inter-
faces 2021,13, 30127.
[384] J. R. Tumbleston, D. Shirvanyants, N. Ermoshkin, R. Janusziewicz,
A. R. Johnson, D. Kelly, K. Chen, R. Pinschmidt, J. P. Rolland, A.
Ermoshkin, E. T. Samulski, J. M. Desimone, Science 2015,347, 1349.
[385] C. Sun, N. Fang, D. M. Wu, X. Zhang, Sens. Actuators, A 2005,121,
113.
[386] Q. Ge, A. H. Sakhaei, H. Lee, C. K. Dunn, N. X. Fang, M. L. Dunn,
Sci. Rep. 2016,6, 31110.
[387] J. J. Schwartz, A. J. Boydston, Nat. Commun. 2019,10, 791.
[388] M. Jamal, A. M. Zarafshar, D. H. Gracias, Nat. Commun. 2011,2,
527.
[389] Z. Zhao, J. Wu, X. Mu, H. Chen, H. J. Qi, D. Fang, Sci. Adv. 2017,3,
e1602326.
[390] Z. Zhao, J. Wu, X. Mu, H. Chen, H. J. Qi, D. Fang, Macromol. Rapid
Commun. 2017,38, 1600625.
[391] Q. Zhang, X. Kuang, S. Weng, Z. Zhao, H. Chen, D. Fang, H. J. Qi,
ACS Appl. Mater. Interfaces 2020,12, 17979.
[392] X. Kuang, J. Wu, K. Chen, Z. Zhao, Z. Ding, F. Hu, D. Fang, H. J. Qi,
Sci. Adv. 2019,5, eaav5790.
[393] B. Jin, J. Liu, Y. Shi, G. Chen, Q. Zhao, S. Yang, Adv. Mater. 34,
2107855.
[394] S. H. Kim, Y. B. Seo, Y. K. Yeon, Y. J. Lee, H. S. Park, M. T. Sultan,
J. M. Lee, J. S. Lee, O. J. Lee, H. Hong, H. Lee, O. Ajiteru, Y. J. Suh,
S.-H. Song, K.-H. Lee, C. H. Park, Biomaterials 2020,260, 120281.
[395] Q. Ge, Z. Chen, J. Cheng, B. Zhang, Y.-F. Zhang, H. Li, X. He, C.
Yuan, J. Liu, S. Magdassi, S. Qu, Sci. Adv. 2021,7, eaba4261.
[396] L. Huang, R. Jiang, J. Wu, J. Song, H. Bai, B. Li, Q. Zhao, T. Xie, Adv.
Mater. 2017,29, 1605390.
[397] J.-T. Miao, M. Ge, S. Peng, J. Zhong, Y. Li, Z. Weng, L. Wu, L. Zheng,
ACS Appl. Mater. Interfaces 2019,11, 40642.
[398] E. Rossegger, R. Höller, D. Reisinger, J. Strasser, M. Fleisch, T.
Griesser, S. Schlögl, Polym. Chem. 2021,12, 639.
[399] B. Zhang, W. Zhang, Z. Zhang, Y.-F. Zhang, H. Hingorani, Z. Liu, J.
Liu, Q. Ge, ACS Appl. Mater. Interfaces 2019,11, 10328.
[400] J. Wu, C. Yuan, Z. Ding, M. Isakov, Y. Mao, T. Wang, M. L. Dunn, H.
J. Qi, Sci. Rep. 2016,6, 24224.
[401] X. Peng, T. Liu, Q. Zhang, C. Shang, Q.-W. Bai, H. Wang, Adv. Funct.
Mater. 2017,27, 1701962.
[402] Z. Ding, O. Weeger, H. J. Qi, M. L. Dunn, Mater. Des. 2018,137, 256.
[403] Z. Ding, C. Yuan, X. Peng, T. Wang, H. J. Qi, M. L. Dunn, Sci. Adv.
2017,3, 1602890.
[404] L. Ionov, Soft Matter 2011,7, 6786.
[405] D. H. Gracias, V. Kavthekar, J. C. Love, K. E. Paul, G. M. Whitesides,
Adv. Mater. 2002,14, 235.
[406] J.-H. Cho, M. D. Keung, N. Verellen, L. Lagae, V. V. Moshchalkov, P.
V. Dorpe, D. H. Gracias, Small 2011,7, 1943.
[407] J. S. Randhawa, T. G. Leong, N. Bassik, B. R. Benson, M. T.
Jochmans, D. H. Gracias, J. Am. Chem. Soc. 2008,130, 17238.
[408] N. Bassik, G. M. Stern, D. H. Gracias, Appl. Phys. Lett. 2009,95,
091901.
[409] D. Martella, S. Nocentini, D. Nuzhdin, C. Parmeggiani, D. S.
Wiersma, Adv. Mater. 2017,29, 1704047.
[410] Y. Guo, J. Zhang, W. Hu, M. T. A. Khan, M. Sitti, Nat. Commun. 2021,
12, 5936.
[411] H. Deng, Y. Dong, J.-W. Su, C. Zhang, Y. Xie, C. Zhang, M. R.
Maschmann, Y. Lin, J. Lin, ACS Appl. Mater. Interfaces 2017,9, 30900.
[412] J. Zhang, Z. Ren, W. Hu, R. H. Soon, I. C. Yasa, Z. Liu, M. Sitti, Sci.
Rob. 2021,6, eabf0112.
[413] G. Villar, A. D. Graham, H. Bayley, Science 2013,340, 48.
[414] T. G. Leong, B. R. Benson, E. K. Call, D. H. Gracias, Small 2008,4,
1605.
[415] W. Xu, T.Li, Z. Qin, Q. Huang, H. Gao, K. Kang, J. Park, M. J. Buehler,
J. B. Khurgin, D. H. Gracias, Nano Lett. 2019,19, 7941.
[416] M. Schenk, S. D. Guest, Origami 2011,5, 291.
[417] Z. Y. Wei, Z. V. Guo, L. Dudte, H. Y. Liang, L. Mahadevan, Phys. Rev.
Lett. 2013,110, 215501.
[418] E. T. Filipov, T. Tachi, G. H. Paulino, in Origami6: II. Technology, Art,
Education (Eds: K. Miura, T. Kawasaki, T. Tachi, R. Uehara, R. J. Lang,
P. Wang-Iverson), American Mathematical Soc., 2015, p.409.
[419] E. T. Filipov, K. Liu, T. Tachi, M. Schenk, G. H. Paulino, Int. J. Solids
Struct. 2017,124, 26.
[420] S. R. Woodruff, E. T. Filipov, Int. J. Solids Struct. 2020,204, 114.
[421] Y. Zhu, E. T. Filipov, J. Mech. Robot. 2020,12.
[422] Y. Zhu, E. T. Filipov, in Volume 8B: 45th Mechanisms and Robotics
Conference (MR), 2021.
[423] A. Ghassaei, E. D. Demaine, N. Gershenfeld, Origami 2018,7, 1151.
[424] K. Liu, G. Paulino, Origami 2018,7, 1167.
[425] R. J. Lang, https://www.langorigami.com/article/tessellatica/
(accessed May, 2022).
[426] E. A. Peraza Hernandez, D. J. Hartl, E. Akleman, D. C. Lagoudas,
Comput. Aided Des. 2016,78, 93.
[427] J. A. Faber, A. F. Arrieta, A. R. Studart, Science 2018,359, 1386.
[428] Z. Zhao, X. Kuang, J. Wu, Q. Zhang, G. H. Paulino, H. J. Qi, D. Fang,
Soft Matter 2018,14, 8051.
[429] H. Fang, S.-C. A. Chu, Y. Xia, K.-W. Wang, Adv. Mater. 2018,30,
1706311.
[430] N. An, M. Li, J. Zhou, Smart Mater. Struct. 2016,25, 11LT02.
[431] G. A. Holzapfel, Nonlinear Solid Mechanics: A Continuum Approach
for Engineering ,Wiley,2000.
[432] L. R. G. Treloar, in The Physics of Rubber Elasticity, 3rd ed., Oxford
University Press, Oxford 2005.
[433] E. M. Arruda, M. C. Boyce, J. Mech. Phys. Solids 1993,41,389.
[434] R. W. Ogden, R. Hill, Proc. R. Soc. London, A 1972,326, 565.
[435] K. K. Westbrook, P. H. Kao, F. Castro, Y. Ding, H. J. Qi, Mech. Mater.
2011,43, 853.
[436] F. Castro, K. K. Westbrook, K. N. Long, R. Shandas, H. J. Qi, Mech.
Time Depend Mater. 2010,14, 219.
[437] H. J. Qi, T. D. Nguyen, F. Castro, C. M. Yakacki, R. Shandas, J. Mech.
Phys. Solids 2008,56, 1730.
[438] Q. Ge, X. Luo, C. B. Iversen, H. B. Nejad, P. T. Mather, M. L. Dunn,
H. J. Qi, Int. J. Solids Struct. 2014,51, 2777.
[439] K. Yu, T. Xie, J. Leng, Y. Ding, H. J. Qi, Soft Matter 2012,8, 5687.
[440] R. Xiao, J. Choi, N. Lakhera, C. M. Yakacki, C. P. Frick, T. D. Nguyen,
J. Mech. Phys. Solids 2013,61, 1612.
[441] Q. Ge, X. Luo, E. D. Rodriguez, X. Zhang, P. T. Mather, M. L. Dunn,
H. J. Qi, J. Mech. Phys. Solids 2012,60, 67.
[442] Y. Liu, K. Gall, M. L. Dunn, A. R. Greenberg, J. Diani, Int. J. Plast.
2006,22, 279.
[443] T. Nguyen, H. J. Qi, F. Castro, K. Long, J. Mech. Phys. Solids 2008,56,
2792.
[444] W. Hong, X. Zhao, J. Zhou, Z. Suo, J. Mech. Phys. Solids 2008,56,
1779.
[445] W. Hong, X. Zhao, Z. Suo, J. Mech. Phys. Solids 2010,58, 558.
[446] F. P. Duda, A. C. Souza, E. Fried, J. Mech. Phys. Solids 2010,58,
515.
[447] S. A. Chester, L. Anand, J. Mech. Phys. Solids 2010,58, 1879.
[448] S. A. Chester, L. Anand, J. Mech. Phys. Solids 2011,59, 1978.
[449] S. Cai, Z. Suo, J. Mech. Phys. Solids 2011,59, 2259.
[450] J. S. Biggins, E. M. Terentjev, M. Warner, Phys. Rev. E 2008,78,
041704.
[451] J. S. Biggins, M. Warner, K. Bhattacharya, Phys.Rev.Lett.2009,103,
037802.
[452] L. Jin, Z. Zeng, Y. Huo, J. Mech. Phys. Solids 2010,58, 1907.
Adv. Mater. 2023, 2302066 © 2023 Wiley-VCH GmbH
2302066 (33 of 35)
15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202302066 by Stanford University, Wiley Online Library on [12/12/2023]. 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 www.advmat.de
[453] J. S. Biggins, M. Warner, K. Bhattacharya, J. Mech. Phys. Solids 2012,
60, 573.
[454] R. Brighenti, C. G. Mcmahan, M. P. Cosma, A. Kotikian, J. A. Lewis,
C. Daraio, Int. J. Solids Struct. 2021,219, 92.
[455] R. Bai, K. Bhattacharya, J. Mech. Phys. Solids 2020,144, 104115.
[456] Y. Zhang, C. Xuan, Y. Jiang, Y. Huo, J. Mech. Phys. Solids 2019,126,
285.
[457] Z. Wang, A. El Hajj Chehade, S. Govindjee, T. D. Nguyen, J. Mech.
Phys. Solids 2022,163, 104829.
[458] R. Zhao, Y. Kim, S. A. Chester, P. Sharma, X. Zhao, J. Mech. Phys.
Solids 2019,124, 244.
[459] D. Mukherjee, M. Rambausek, K. Danas, J. Mech. Phys. Solids 2021,
151, 104361.
[460] L. Wang, Y. Kim, C. F. Guo, X. Zhao, J. Mech. Phys. Solids 2020,142,
104045.
[461] D. Yan, A. Abbasi, P. M. Reis, Int. J. Solids Struct. 2021, 111319.
[462] T. G. Sano, M. Pezzulla, P. M. Reis, J. Mech. Phys. Solids 2022,160,
104739.
[463] W. Chen, L. Wang, J. Appl. Mech. 2020,87.
[464] R. Long, H. J. Qi, M. L. Dunn, Soft Matter 2013,9, 4083.
[465] R. Long, H. J. Qi, M. L. Dunn, J. Mech. Phys. Solids 2013,61, 2212.
[466] K. N. Long, T. F. Scott, H. J. Qi, C. N. Bowman, M. L. Dunn, J. Mech.
Phys. Solids 2009,57, 1103.
[467] K. N. Long, J. Mech. Phys. Solids 2014,63, 386.
[468] F. J. Vernerey, R. Long, R. Brighenti, J. Mech. Phys. Solids 2017,107,
1.
[469] F. J. Vernerey, J. Mech. Phys. Solids 2018,115, 230.
[470] J. Ma, X. Mu, C. N. Bowman, Y. Sun, M. L. Dunn, H. J. Qi, D. Fang,
J. Mech. Phys. Solids 2014,70, 84.
[471] T. D. Nguyen, Polym. Rev. 2013,53, 130.
[472] K. N. Long, M. L. Dunn, H. J. Qi, Int. J. Plast. 2010,26, 603.
[473] P. J. Flory, in Principles of Polymer Chemistry, Cornell University Press,
New York 1953.
[474] M. L. Huggins, J. Chem. Phys. 1941,9, 440.
[475] P. J. Flory, J. Chem. Phys. 1942,10, 51.
[476] W. Hong, Z. Liu, Z. Suo, Int. J. Solids Struct. 2009,46, 3282.
[477] S. A. Chester, C. V. Di Leo, L. Anand, Int. J. Solids Struct. 2015,52,1.
[478] M. J. Stephen, J. P. Straley, Rev. Mod. Phy. 1974,46, 617.
[479] P. Bladon, E. M. Terentjev, M. Warner, Phys.Rev.E1993,47, R3838.
[480] G. C. Verwey, M. Warner, Macromolecules 1997,30, 4189.
[481] A. Dorfmann, R. W. Ogden, Eur. J. Mech. A Solids 2003,22, 497.
[482] K. Danas, S. V. Kankanala, N. Triantafyllidis, J. Mech. Phys. Solids
2012,60, 120.
[483] L. Dorfmann, R. W. Ogden, in Nonlinear Theory of Electroelastic and
Magnetoelastic Interactions, 1st ed., Springer, New York, NY 2014.
[484] L. Wang, D. Zheng, P. Harker, A. B. Patel, C. F. Guo, X. Zhao, Proc.
Natl. Acad. Sci. USA 2021,118, e2021922118.
[485] D. Yan, M. Pezzulla, L. Cruveiller, A. Abbasi, P. M. Reis, Nat. Com-
mun. 2021,12, 2831.
[486] Y. Mao, F. Chen, S. Hou, H. J. Qi, K. Yu, J. Mech. Phys. Solids 2019,
127, 239.
[487] K. N. Long, M. L. Dunn, T. F. Scott, L. P. Turpin, H. J. Qi, J. Appl. Phys.
2010,107, 053519.
[488] K. N. Long, T. F. Scott, M. L. Dunn, H. J. Qi, Int. J. Solids Struct. 2011,
48, 2089.
[489] X. Mu, N. Sowan, J. A. Tumbic, C. N. Bowman, P. T. Mather, H. J. Qi,
Soft Matter 2015,11, 2673.
[490] X. Sun, H. Wu, R. Long, Soft Matter 2016,12, 8847.
[491] Q. Guo, R. Long, in Mechanics of Polymer Networks with Dynamic
Bonds, in Self-Healing and Self-Recovering Hydrogels (Eds: C. Creton,
O. Okay), Springer, Cham 2020, p. 127.
[492] C.-Y. Hui, F. Cui, A. Zehnder, F. J. Vernerey, Proc. R. Soc. A 2021,477,
20210608.
[493] C. Luo, Z. Lei, Y. Mao, X. Shi, W. Zhang, K. Yu, Macromolecules 2018,
51, 9825.
[494] S. Akbari, A. H. Sakhaei, K. Kowsari, B. Yang, A. Serjouei, Z.
Yua n fang , Q . Ge, Smart Mater. Struct. 2018,27, 065027.
[495] Z. Zhao, X. Kuang, C. Yuan, H. J. Qi, D. Fang, ACS Appl. Mater. In-
terfaces 2018,10, 19932.
[496] Y. Liu, Y. Li, G. Yang, X. Zheng, S. Zhou, ACS Appl. Mater. Interfaces
2015,7, 4118.
[497] S. Weng, X. Kuang, Q. Zhang, C. M. Hamel, D. J. Roach, N. Hu, H.
J. Qi, ACS Appl. Mater. Interfaces 2021,13, 12797.
[498] S. H. Aswathy, U. Narendrakumar, I. Manjubala, Heliyon 2020,6,
e03719.
[499] Y. Tang, Y. Chi, J. Sun, T.-H. Huang, O. H. Maghsoudi, A. Spence, J.
Zhao, H. Su, J. Yin, Sci. Adv. 2020,6, eaaz6912.
[500] M. Ha, G. S. Cañón Bermúdez, J. A.-C. Liu, E. S. Oliveros Mata, B.
A. Evans, J. B. Tracy, D. Makarov, Adv. Mater. 2021,33, 2008751.
[501] C. Becker, B. Bao, D. D. Karnaushenko, V. K. Bandari, B. Rivkin, Z. Li,
M. Faghih, D. Karnaushenko, O. G. Schmidt, Nat. Commun. 2022,
13, 2121.
Sophie Leanza is currently an undergraduate Chemical Engineering student at The Ohio State Univer-
sity and a research assistant for Stanford University.She will pursue a Ph.D. in Mechanical Engineering
at Stanford University following her undergraduate studies. Her research interests include stimuli-
responsive materials, soft robotics, and the shape morphing of structures.
Adv. Mater. 2023, 2302066 © 2023 Wiley-VCH GmbH
2302066 (34 of 35)
15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202302066 by Stanford University, Wiley Online Library on [12/12/2023]. 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 www.advmat.de
Shuai Wu is currently a Ph.D. candidate in Mechanical Engineering at Stanford University. He received
his B.S. degree in Engineering Mechanics from Southwest Jiaotong University and M.S. degree in
Mechanical Engineering from University of California, San Diego. His research interests focus on
reconfigurable soft materials and structures and soft robotics.
Xiaohao Sun is a postdoctoral fellow in the School of Mechanical Engineering at Georgia Institute of
Technology.Prior to that, he was a postdoctoral associate in the Mechanical Engineering Department
at University of Colorado Boulder. He received his B.S. degree in theoretical and applied mechanics
from the University of Science and Technology of China in 2014, and Ph.D. degree in solid mechan-
ics from the same university in 2019. His research interests include mechanics of soft materials and
modeling and design for 4D printing.
H. Jerry Qi is a Professor of Mechanical Engineering at Georgia Institute of Technology and is the site
director of NSF Industry–University Cooperative Research Center on Science of Heterogeneous Ad-
ditive Printing of 3D Materials (SHAP3D). His research focuses on developing fundamental under-
standing of multi-field properties of soft active materials, including shape memory polymers, light
activated polymers, liquid crystal elastomers, and vitrimers. He has been working on integrating active
materials with 3D printing and developing new materials and new methods in polymer 3D printing for
applications in morphing structures, metamaterials, origami, and soft robotics.
Ruike Renee Zhao is an Assistant Professor of Mechanical Engineering at Stanford University where
she directs the Soft Intelligent Materials Laboratory.Renee’s research concerns the development
of stimuli-responsive soft composites for multifunctional robotic systems with integrated shape-
changing, assembling, sensing, and navigation. By combining mechanics, polymer engineering, and
advanced material manufacturing techniques, the functional soft composites enable applications
in soft robotics, miniaturized biomedical devices, flexible electronics, and deployable and morphing
structures.
Adv. Mater. 2023, 2302066 © 2023 Wiley-VCH GmbH
2302066 (35 of 35)
15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202302066 by Stanford University, Wiley Online Library on [12/12/2023]. 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
... 4D printing is a common method for demonstrating the morphing of simple origami structures by integrating functional materials and additive manufacturing. [42,50,51] Embedding entropy, for example, internal stress, in the 3D printing of smart materials, is a typical strategy in 4D printing. [52][53][54] However, most 4D printing techniques are limited with respect to the synthesis of strong, morphable, and 3D curvilinear structures. ...
... Traditional active origami structures often depend on hinge flexibility, achieved by one of three methods: reducing the hinge's thickness, cutting the hinge, or using materials with reduced stiffness at the hinge. [51] This method tends to focus stress on the weakened hinges. In contrast, our panel deformation-driven method enables the use of hinges with thicker and stiffer materials, which results in markedly enhanced structural stiffness. ...
Article
Full-text available
Morphing origami has numerous potential engineering applications owing to its intrinsic morphing features from 2D planes to 3D surfaces. However, the current 1D hinge deformation‐driven transformation of foldable origami with rigid or slightly deformable panels cannot achieve a 3D complex and large curvilinear morphing. Moreover, most active origami structures use thin hinges with soft materials on their creases, thus resulting in a lower load capability. This study proposes a novel origami morphing method demonstrating large free‐form surface morphing, such as Euclidean to non‐Euclidean surface morphing with shape‐locking. Tensorial anisotropic stress in origami panels is embedded during the extrusion‐based 3D printing of shape memory polymers. The extrusion‐based 3D printing of isotropic SMPs can produce tensorial anisotropic stress in origami panels during fabrication, which can realize significant non‐Euclidean surface morphing with multiple deformation modes. The connecting topology of the origami unit cells influences the global morphing behavior owing to the interaction of the deformation of adjacent panels. Non‐Euclidean morphing integrated with 4D printing can provide multimodal shape locking at material and structural levels.
... Once external stimuli are removed, LCEs fully recover to their initial shape, showing good reversibility. However, the actuation and recovery of the LCE are relatively slow compared to DEAs and conventional machines, significantly restricting their applications regardless of their large deformation, high work density, and easy actuating methods ( Figure 2) [25][26][27][28][29][30]. Thus, it is emerging to develop strategies to produce rapid actuation in LCE materials. ...
Article
Liquid crystal elastomer (LCE) is one kind of soft actuating material capable of producing large and reversible actuation strain, versatile and programmable actuation modes, and high work density, which can be widely exploited for nextgeneration soft robots. However, the slow response speed and low power density in LCE-based actuators remain a challenge, limiting their practical applications. Researchers have been considering how to improve these performances. In this review, we discuss the fundamentals of the LCEs and emphasize the fast actuation strategies developed in recent years. Firstly, we introduce conventional preparation strategies. Then, we describe typical actuation mechanisms of LCEs, discussing their features and limitations. Subsequently, we summarize several possible approaches as case studies to enhance the actuation performance of LCEs, including reducing physical sizes, introducing active heating-cooling mechanisms, utilizing mechanical instability, and developing dielectric LCEs. Finally, we discuss the future research opportunities and challenges for rapid actuation of LCEs.
Article
Full-text available
Origami, the art of paper folding, has emerged as a versatile technique for crafting intricate 3D structures from 2D sheets. Combined with the magnetic actuation, origami paper becomes the building blocks for cost‐effective, wirelessly controllable magnetic robots. Herein, a biodegradable magnetic paper with excellent formability and recyclability is developed, facilitating its convenient utilization and disposal. The programable magnetic paper, fabricated with specific magnetization and crease patterns, enables the transformation of 2D sheets into predetermined 3D structures. Leveraging the lightweight and pliable nature of paper‐based materials, exceptional control of origami robots with fast response is demonstrated, enabling a wide range of locomotion. Furthermore, the paper‐based approach enables the incorporation of electronic functionality into magnetic actuators. By introducing conductive nanoparticles into magnetic paper, an electrically conductive substance is created. Constructing electronic circuits and integrating electronic components onto the paper‐based printed circuit board platform enables the repairing of broken circuits inside complicated equipment and optical sensing of surrounding environments in conjunction with locomotive robots. The origami robots have a huge potential to be facilitated in diverse fields with various functions, demonstrating complex locomotion, and integrating chemical, optical, thermal, and mechanical sensors for monitoring environmental conditions in hard‐to‐reach locations. The array of possibilities holds significant promise for the widespread application of these origami magnetic robots across a diverse spectrum of research fields in soft robotics.
Article
Full-text available
This study presents a concept for a straightforward method to enhance the actuation performances of magneto‐active elastomer membranes. The concept is based on a characteristic magnetization pattern and offers a solution to two major difficulties in the actuation of thin and mechanically soft magnetic actuators: the localization of actuation forces and the self‐demagnetization. After the magnetization process, the membrane presents two regions with an oppositely oriented out‐of‐plane magnetization. The magnetized regions are separated by a transition zone which is called magnetic pole transition. Experimental investigations reveal a high magnetic flux density near the pole transition—even in the center of bidirectionally magnetized membranes—whereas the magnetic flux density of a uniformly magnetized membrane decreases toward the center. In additional experiments, membranes with both magnetization patterns are actuated by stiff permanent magnets. The resulting out‐of‐plane displacement of the bidirectionally magnetized membrane exceeds the displacement of the unidirectionally magnetized membrane by far. The investigations demonstrate that this enhancement stems from the presence of the magnetic pole transition. All experiments are reproduced using magnetic and magneto‐mechanical numerical models; a good accordance between the results is achieved.
Article
Full-text available
Soft linear actuators have strong deformation ability and good environmental adaptability, which have been widely used in soft robot design. However, little work has focused on designing soft linear actuators with balanced performances, featuring fast driving speed, large output displacement, lightweight, and miniaturization. Herein, we present a novel soft linear actuator design based on the Kresling origami structure. By driving the Kresling tubes with a servo motor, the soft linear actuator has good compliance and strong environmental adaptability and can achieve fast driving speed, large driving force, and high control precision comparable to the traditional electrical motor. The analytical models of the Kresling tubes and the whole actuator are respectively derived to analyze the mechanical properties, determine the optimal geometry of the Kresling tube and evaluate the driving performance of the whole actuator. The actuator prototype is fabricated by 3D printing, and the actual driving performance is tested. It is shown that the prototype can achieve a maximum output displacement of 18.9 mm without payload or 16mm under a payload of 30 N. Finally, as a case study, the soft linear actuator is applied to a crawling robot, where the maximum moving speed of 28 mm/s is reached.
Article
Interlocking assemblies have been explored to address large assembly and complex parts and are now integral to additive manufacturing (AM) for creating objects with dissimilar materials and multiple properties. 4D printing technology, which combines smart materials (SMs) with AM, aligns with this approach by enabling the creation of objects that can change shape or properties when exposed to external stimuli. As 4D printing trends towards strategically arranging active and passive materials for improved control and performance, challenges arise due to the limited compatibility of existing 3D printers with the required SM properties. To address this issue, interlocking blocks of dissimilar materials can be printed and then assembled into a desired shape. This work aims to generalize the applicability of the interlocking block assembly approach. This will be achieved by improving the deformation uniformity in a 4D multi-material interlocked assembly. This paper also addresses limitations that can occur due to the interfaces between interlocking blocks, such as lack of deformation and contact continuity. Thus, it will be a question of customizing the shapes of the blocks in the early stages of assembly generation, considering SMs and their potential transformations. Finally, this approach is illustrated with an example, shedding light on the practical implications.
Article
Full-text available
The last few decades have witnessed unprecedented convergence between humans and machines that closely operate around the human body. Despite these advances, traditional machines made of hard, dry and abiotic materials are substantially dissimilar to soft, wet and living biological tissues. This dissimilarity results in severe limitations for long-term, reliable and highly efficient interfacing between humans and machines. To bridge this gap, hydrogels have emerged as an ideal material candidate for interfacing between humans and machines owing to their mechanical and chemical similarities to biological tissues and the versatility and flexibility in designing their properties. In this Review, we provide a comprehensive summary of functional modes, design principles, and current and future applications for hydrogel interfaces towards merging humans and machines. Hydrogels are one of the most promising materials to bridge the stark disparities between traditional machines and biological tissues for successful interfacing between humans and machines. This Review discusses the functional modes, design principles, and current and future applications of hydrogel interfaces for the merging of humans and machines.
Article
Full-text available
This study reports fabrication of highly porous electrospun self‐folding bilayers, which fold into tubular structures with excellent mechanical stability, allowing them to be easily manipulated and handled. We fabricated and compared two kinds of bilayers based on FDI‐approved biocompatible and biodegradable soft (PCL, polycaprolactone) and hard (PHB, poly‐hydroxybutyrate) thermoplastic polymers. Multi‐scroll structures with tunable diameter in aqueous media were obtained after the shape transformation of the bilayer, where PCL‐based bilayer rolled longitudinally and PHB‐based one rolled transversely with respect to fiber direction. A combination of higher elastic modulus and transverse orientation of fibers with respect to rolling direction allowed precise temporal control of shape transformation of PHB‐bilayer – stress produced by swollen methacrylated hyaluronic acid (HA‐MA) did not relax with time and folding was not affected by the fact that bilayer was fixed in unfolded state in cell culture medium for more than 1 h. This property of PHB‐bilayer allowed cell culturing without negative effect on its shape transformation ability. Moreover, PHB‐based tubular structure demonstrated superior mechanical stability compared to PCL‐based ones and did not collapse during manipulations that happened to PCL‐based one. Additionally, PHB/HA‐MA bilayers showed superior biocompatibility, degradability, and long‐term stability compared to PCL/HA‐MA. After 14 days of cultivation, our vascular grafts showed high stability in cultivation, and between 50–70% of the multi‐scroll area was fully covered. All these make bilayer of transversely aligned hard thermoplastic polymer fibers (such as PHB) and swelling hydrophilic polymer fibers (such as HA‐MA) more suitable for the fabrication of blood vessel replacement in comparison to widely used PCL‐based bilayers. This article is protected by copyright. All rights reserved
Article
Full-text available
Liquid crystal elastomers (LCE) are appealing candidates among active materials for 4D printing, due to their reversible, programmable, and rapid actuation capabilities. Recent progress has been made on direct ink writing (DIW) or digital light processing (DLP) to print LCEs with certain actuation. However, it remains a challenge to achieve complicated structures, such as spatial lattices with large actuation, due to the limitation of printing LCEs on the build platform or the previous layer. Herein, we propose a novel method to 4D print freestanding LCEs on‐the‐fly by using laser‐assisted DIW with an actuation strain up to ‐40%. This process is further hybridized with the DLP method for optional structural or removable supports to create active 3D architectures in a one‐step additive process. We demonstrate that various objects, including hybrid active lattices, active tensegrity, an actuator with tunable stability, and 3D spatial LCE lattices, can be additively fabricated. The combination of DIW‐printed functionally freestanding LCEs with the DLP‐printed supporting structures thus provides new design freedom and fabrication capability for applications including soft robotics, smart structures, active metamaterials, and smart wearable devices. This article is protected by copyright. All rights reserved
Book
This book provides a critical review of the equilibrium elastic properties of rubber, together with the kinetic-theory background. It is suitable for the non-specialist and the emphasis is on the physical reality embodied in the mathematical formulations. Polymer science had developed greatly since the second edition of this text in 1958, and the two main advances – the refinements of the network theory and associated thermodynamic analysis, and the development of the phenomenological or non-molecular approach to the subject – are both reflected in the structure of this third edition.
Book
This new edition of this classic text on condensed matter physics includes the great advances that have taken place since its first publication in 1974. New chapters describe the main types and properties of liquid crystals in terms of the new phases discovered since the middle of the 1970's, and advances in the understanding of local order and the nature of the isotropic to nematic transition. There is an extensive discussion of the symmetry, and macroscopic and dynamic properties of smectics and columnar phases, and their defects, illustrated with numerous descriptions of experimental arrangements. The final chapter is devoted to phase transitions in smectics, including the celebrated analogy between Smectic A and superconductors. Throughout the book there is an emphasis on order-of-magnitude considerations. Its topicality and breadth of coverage will ensure that The Physics of Liquid Crystals remains an indispensable guide for students and researchers alike.
Article
Shape transformation and motion guidance are emerging research hotspots of mechanical metamaterials. In this case, the key issue is how to improve the programmability and reconfigurability of metamaterials. The magnetically driven method enables materials to accomplish remote, fast, and reversible deformation, so it is desired for improving the programmability and reconfigurability of metamaterials. However, conventional magnetically driven materials are often pure elastomer materials. Their magnetic programming method is single, and their overall shape is unchangeable after fabrication, which limits their programmability and reconfigurability. Herein, this article proposes a kind of magnetically driven, programmable, and reconfigurable modular mechanical metamaterial based on origami and kirigami design mechanisms. The motion and deformation were designed to follow the predefined creases and incisions that could be transformed into each other. This metamaterial enabled more discrete motion and force transmission and integrated the fold of origami, the rotation of kirigami, and the fold guided by cuts. Such designs laid the foundation for complex, three-dimensional structures which could be quickly reassembled and constructed to deal with complex situations. This paper also demonstrated applications of this metamaterial in information storage and manifestation, mechanical logic computing, reconfigurable robotics, deployable mechanisms, and so on. The results indicated that the high programmability and reconfigurability expanded the application potential of the metamaterial for broader needs.
Article
Hexagonal ring origami is a type of foldable structure that has impressive packing abilities and can be tessellated into two-dimensional or three-dimensional surfaces without any gap or overlap. It can be folded under bending or twisting loads into a peach core-shaped configuration with only 10.6% of its initial area. However, in applications of large-scale foldable structures, folding by bending or twisting is usually technically difficult. Here, we propose strategies to facilitate easy snap-folding of the hexagonal ring by a simple point load or localized twist or squeeze. This is enabled by two geometric modifications made to the hexagonal ring: introducing residual strain and creating pre-twisted edges. By combining theoretical modeling, finite element simulations, and experiments, we systematically investigate the snap-folding behaviors of modified hexagonal rings with residual strain and pre-twisted edges. It is found that the geometric modifications promote easy snap-folding of the hexagonal ring by different mechanisms: introducing residual strain can significantly decrease the energy barrier and thus reduce the required moment to snap-fold the ring, while creating pre-twisted edges allows for easy out-of-plane deformation which is a necessary condition for a ring to fold. Combining the two methods further enables the snap-folding of the hexagonal ring by a point load or localized twist or squeeze. To demonstrate the easy folding of large assemblies of the modified rings, we construct dome and pyramid assemblies that can be snap-folded from their initial three-dimensional states to significantly lower-volume final states by a simple compression at the rings’ corners. We envision that the proposed geometric modification strategies can provide a new perspective on the rational design of easy-to-fold ring origami-based foldable functional structures with extremely high packing ratios.
Article
Functional structures with reversible shape-morphing and color-changing capabilities are promising for applications including soft robotics and biomimetic camouflage devices. Despite extensive studies, there are few reports on achieving both reversible shape-switching and color-changing capabilities within one structure. Here, we report a facile and versatile strategy to realize such capabilities via spatially programmed liquid crystal elastomer (LCE) structures incorporated with thermochromic dyes. By coupling the shape-changing behavior of LCEs resulting from the nematic-to-isotropic transition of liquid crystals with the color-changing thermochromic dyes, 3D thermochromic LCE structures change their shapes and colors simultaneously, which are controlled by the nematic-isotropic transition temperature of LCEs and the critical color-changing temperature of dyes, respectively. Demonstrations, including the simulated blooming process of a resembled flower, the camouflage behavior of a "butterfly"/"chameleon" robot in response to environmental changes, and the underwater camouflage of an "octopus" robot, highlight the reliability of this strategy. Furthermore, integrating micro-ferromagnetic particles into the "octopus" thermochromic LCE robot allows it to respond to thermal-magnetic dual stimuli for "adaptive" motion and diverse biomimetic motion modes, including swimming, rolling, rotating, and crawling, accompanied by color-changing behaviors for camouflage. The reversibly reconfigurable and color-changing thermochromic LCE structures are promising for applications including soft camouflage robots and multifunctional biomimetic devices.