ArticlePDF Available

MEMS actuators for biomedical applications: A review

Authors:

Abstract and Figures

Micro-electromechanical-system (MEMS) based actuators, which transduce certain domains of energy into mechanical movements in the microscopic scale, are increasingly contributing to the areas of biomedical engineering and healthcare applications. They are enabling new functionalities in biomedical devices through their unique miniaturized features. An effective selection of a particular actuator, among a wide range of actuator types available in the MEMS field, needs to be made through the assessment of many factors involved in both the actuator itself and the target application. This paper presents an overview of the state-of-the-art MEMS actuators that have been developed for biomedical applications. The actuation methods, working principle, and imperative features of these actuators are discussed along with their specific applications. An emphasis of this review is placed on temperature-responsive, electromagnetic, piezoelectric, and fluid-driven actuators towards various application areas including lab-on-a-chip, drug delivery systems, cardiac devices and surgical tools. It also highlights the key issues of MEMS actuators in light of biomedical applications.
Content may be subject to copyright.
The published version is available at https://doi.org/10.1088/1361-6439/ab8832
1
MEMS Actuators for Biomedical Applications: A Review
Farah Afiqa Mohd Ghazali1, Md. Nazibul Hasan1, Tariq Rehman1,2, Marwan Nafea3,
Mohamed Sultan Mohamed Ali1, and Kenichi Takahata4
1 School of Electrical Engineering, Universiti Teknologi Malaysia, 81310 UTM Johor Bahru, Johor, Malaysia
2 Department of Electronic Engineering, Faculty of Electrical and Computer Engineering, NED University of
Engineering and Technology, Karachi 75270, Pakistan
3 Department of Electrical and Electronic Engineering, University of Nottingham Malaysia, Jalan Broga,
43500, Semenyih, Selangor, Malaysia
4 Department of Electrical and Computer Engineering, University of British Columbia, Vancouver, B.C.
V6T1Z4, Canada
E-mail: takahata@ece.ubc.ca and sultan_ali@fke.utm.my
Abstract
Micro-electromechanical-system (MEMS) based actuators, which transduce certain domains
of energy into mechanical movements in the microscopic scale, are increasingly contributing
to the areas of biomedical engineering and healthcare applications. They are enabling new
functionalities in biomedical devices through their unique miniaturized features. An effective
selection of a particular actuator, among a wide range of actuator types available in the
MEMS field, requires to be made through assessment of many factors involved in both the
actuator itself and a target application. This paper presents an overview of the state-of-the-art
MEMS actuators that have been developed for biomedical applications. The actuation
methods, working principle, and imperative features of these actuators are discussed along
with their specific applications. An emphasis of this review is placed on temperature-
responsive, electromagnetic, piezoelectric, and fluid-driven actuators towards various
application areas including lab-on-a-chip, drug delivery systems, cardiac devices, and
surgical tools. It also highlights the key issues of MEMS actuators in light of biomedical
applications.
Keywords: MEMS, actuators, biomedical devices, lab-on-a-chip, smart implants, surgical
devices
The published version is available at https://doi.org/10.1088/1361-6439/ab8832
2
1. Introduction
The rapid development of micro-electromechanical-system (MEMS) technologies has
increasingly provided means to miniaturize and advance various biomedical devices and
bioMEMS [1-4]. The applications of these MEMS-based devices include cardiac devices [5,
6], microneedles [7, 8], lab-on-a-chip devices for fast chemical/biological analysis [9-11],
microsurgical robots [12-16], and in-vivo drug delivery systems for drug release with
precision dosage and timing control [17-20]. MEMS actuators are widely used to realize
these types of devices and enable accurate control of them [21]. Serving as core architectural
elements, MEMS actuators have emerged as a promising technology that plays a vital role in
enabling a wide range of biomedical devices. Among existing MEMS actuators, those with
thermoresponsive [22], electromagnetic [23], piezoelectric [24], thermopneumatic [25], and
pneumatic [26] mechanisms have been some of the representative types widely used for
biomedical applications. Each of these actuator types possesses attractive features. For
instance, shape memory alloys (SMAs), a type of smart materials that respond to temperature,
offer high work density, large actuation force and displacement, simple structural design,
resistance to corrosion, and biocompatibility [27-29]. Electromagnetic actuators generally
provide large displacement, fast dynamic response, and an ability of low-voltage and remote
actuation [30-32]. Piezoelectric actuators are often used in ultra-precision and high-speed
applications due to their ability of nano-scale actuation, quick response, and self-locking at
power-off state [33-36]. Pneumatic microactuators are well-known for simple structure, high
flexibility, high force per unit volume, high energy density, and low cost [37-41].
The capabilities of MEMS actuators are continuously growing with a great promise for
diverse future applications. As those actuators exhibit different characteristics and
shortcomings, however, a particular type should be wisely selected and applied for a targeted
biomedical device while assessing the requirements involved in the device and its
environment. In this paper, the working principles, designs, characteristics, and their key
applications of MEMS actuators are comprehensively discussed with an aim to aid further
development of bioMEMS and other biomedical microdevices functionalized by the
actuators. This review is structured as follows: The working principles of thermoresponsive,
electromagnetic, piezoelectric, and fluid-driven microactuators are discussed in Section 2.
Section 3 presents critical applications of these actuators, including lab-on-a-chip, drug
delivery systems, cardiac devices, and surgical and endoscopic tools. The review is
The published version is available at https://doi.org/10.1088/1361-6439/ab8832
3
concluded with a discussion of major factors toward enabling elevated performance of these
actuators in Section 4.
2. Types of Biomedical MEMS Actuators
2.1. Thermoresponsive Actuators
SMAs, shape memory polymers (SMPs), and certain types of hydrogels are classified as
smart materials that have an ability of shape recovery when triggered by an environmental
stimulus. They commonly respond to heat, whereas specific responsive hydrogels also trigger
with others such as radiation, moisture, pH level, and magnetic and electric fields [42-47].
This section reviews these thermoresponsive smart materials regarding their phase transition
modes and characteristics that allow them to work as actuators in the micro domain.
2.1.1 SMA
The actuation of SMAs is based on the principle of a shape-memory effect called martensitic-
austenitic transformation. When a SMA is in its martensite phase, the alloy is in the form of
monoclinic crystals, which makes it more flexible and hence more easily deformed.
Following the deformation of the material’s crystalline orientation, cubic crystals are
constructed within the molecular arrangement, while the material becomes rigid and hard to
deform above the austenite temperature upon heating. When a SMA is cooled in the absence
of a load, the materials crystal structure follows twinned martensite. During this phase, the
SMA can be deformed by applying an external force or by employing a bias spring to achieve
reversible motion. The changes in the crystalline state of SMA are illustrated in Figure 1a
[48]. There are several phase transformation temperatures that must be considered when
selecting a SMA with respect to its applications. During the shape recovery process, the
transformation from the martensite cold state to the austenite hot state begins at the austenite
starting temperature and ends at the austenite finishing temperature. Meanwhile, the
transformation from the hot austenite phase to the cold martensite phase begins at the
martensite starting temperature and ends at the martensite finishing temperature. The SMA
typically consists of a few elements, and the composition level among these elements
determines the transformation temperature. In other words, the elemental composition can be
adjusted to achieve a specific transformation temperature depending on the application.
SMA actuators in the MEMS area are typically fabricated in a form of patterned thin
film or bulk-micromachined structures [49-59]. They possess general attractive attributes
The published version is available at https://doi.org/10.1088/1361-6439/ab8832
4
including large displacement, large force, high mechanical robustness, and corrosion-resistant
[60-66]. The NiTi alloy known as Nitinol is one of the most widely used SMA materials for
biomedical applications owing to its high biocompatibility that facilitates the application for
implantable devices such as surgical tools, cardiac devices, and drug delivery systems [22,
56, 67-71]. General disadvantages of SMA actuators lie in relatively slow temporal response
as well as high power consumption when actuated with self-heating by passing an electrical
current to the material.
2.1.2 SMP
The SMPs have gained significant interest in biomedical applications due to its general
features such as structural flexibility, large strains, low density, tunable transition
temperature, and biodegradable properties [72, 73]. These features make them suitable for
applications in endovascular and drug delivery devices [74, 75]. The thermoresponsive SMP
exhibits a shape-memory effect based on the polymer’s dual-segment system comprised of
cross-links and switching segments. The cross-links determine the permanent shape of the
polymer whereas the switching segments coupled with transition temperature fix the
temporary shape. The SMP is stiff when its temperature is below the transition temperature,
whereas heating it over the transition temperature makes it relatively soft. For shape setting,
an external force must be applied to an SMP while it is heated above the transition
temperature. This step causes the switching segments to fix the molecular chain positions.
Afterwards, the SMP is cooled while removing the external force to result in a memorized
Figure 1. Phase transformations of shape-memory materials. (a) Changes in the crystalline
orientation of SMA at different phases. Reproduced with permission [48]. Copyright 2016,
Elsevier. (b) Shape recovery process in SMP. Reproduced with permission [76]. Copyright
2016, Wiley-VCH.
The published version is available at https://doi.org/10.1088/1361-6439/ab8832
5
shape for the polymer. Applying heat to the SMP induces recovery of the memorized shape
through the shape memory effect as illustrated in Figure 1b [76]. Although this actuator
possesses the aforementioned beneficial properties, SMPs often suffer from slow response
and low recovery stress.
2.1.3 Temperature-Sensitive Hydrogels
Hydrogels are three-dimensional polymeric networks with hydrophilic structures that allow
the absorption of a large amount of aqueous solution in the networks [77]. Depending on the
type of cross-linking between polymers, some of them display mass reversible changes in
response to physical or chemical stimulus [77]. Poly(N-isopropyl acrylamide), or PNIPAM in
short, is a thermoresponsive hydrogel that changes its size at a phase transition temperature
called the lower critical solution temperature (LCST) in the solution [78]. When temperature
of PNIPAM hydrogel is raised above the LCST, the material shrinks by releasing the uptake
solution, whereas reducing the temperature reverses the process [79]. Different material
compositions of PNIPAM can be used to modify its LCST level to tailor it to a specific
application [80]. Besides intrinsic phase transition behavior, the hydrogels also possess
distinct attributes such as tunable mechanical and degradation features, sensitivity towards
stimuli, and ability to conjugate with hydrophilic and hydrophobic therapeutic compounds.
Additionally, PNINAM can be synthesized to be ultraviolet-light sensitive in its
polymerization, which enables precise patterning and complex structure formation of the
polymer through a photolithographic process [81]. These features have promoted the
application of PNIPAM for biomedical devices, such as microvalves in drug delivery systems
as well as encapsulation and delivery of cells [79, 82-90]. In spite of many advantages,
thermoresponsive hydrogels inherently suffer from relatively slow temporal responses similar
to SMA and SMP, and may pose leakage of the solution through the material.
2.2 Electromagnetic Actuators
Electromagnetic actuators generally employ the interaction of one or more magnetic
structures with the magnetic field (B) produced by a current-carrying circuit [91]. A common
configuration of these actuators consists of a coil and a ferromagnetic movable structure
placed in the field produced by the coil as illustrated in Figure 2 with a suspended cantilever
beam being the movable magnetic structure [92, 93]. When the driving current, i, is passed
through the coil, it produces B defined by Biot-Savart law [94, 95] as:
The published version is available at https://doi.org/10.1088/1361-6439/ab8832
6
(1)
where μ0, μr, Ni, and l are the permeability of free space, the relative permeability of the
material, the number of the coil’s turns, and the length of the coil, respectively. The
interaction with B induces an attractive force, F, acting on the cantilever beam to cause a
displacement, X, at the beam’s free end, which can be expressed as [96]:
(2)
where L, E, w, and t are the length of the beam, the Young’s modulus of the material, the
width of the beam, and the thickness of the beam, respectively. This type of actuators has
been used in various MEMS applications given its advantages such as simple drive mode,
high field energy density, fast response time, and large deflection that are attainable with low
input voltages [31, 32]. Their applications extend to micro positioning systems [97],
micromirrors [98, 99], microgrippers [100], and microfluidics [23, 101] for micropumps
[102, 103] and microvalves [104]. Electromagnetic actuators also exhibit common
disadvantages, e.g., volumetric scaling of produced electromagnetic forces that rapidly drop
Figure 2. Schematic diagram on the working mechanism of an electromagnetic actuator
under (a) the off state without current and (b) the on state with a driving current fed to the
solenoidal coil
creating a magnetic field to displace the ferromagnetic movable
microstructure.
The published version is available at https://doi.org/10.1088/1361-6439/ab8832
7
as the device size shrinks, high power dissipation for driving coils, and parasitic loss at high
frequency [105], which should be taken into account in the design of application device.
2.3 Piezoelectric Actuators
Piezoelectric actuators have been widely adopted in the fields of ultra-precision engineering
and microactuation owing to its advantageous features such as fast response, high
displacement resolution, high efficiency, compact structure, and immunity to magnetic field
[34, 106-108]. The operation of the actuators relies on the converse piezoelectric effect of a
piezoelectric crystal to induce strain by applying an electric potential to the crystalline
material [109]. The converse piezoelectric effect can be theoretically described with the
following relationships [110]:
(3)
(4)
where S, E, sE, T, D, d, and ε are the strain, the electric field, the compliance with zero field,
the surface stress, the charge displacement, the piezoelectric strain coefficient, and the
dielectric constant of a piezoelectric material, respectively. The performance of this type of
actuators largely depends on the crystal structure of a piezoelectric material where d acts as a
medium for the transduction mechanism. Given the orientations of polarization and electric
field (P and E, respectively), three different modes, i.e., longitudinal mode (d33), transversal
mode (d31), and shear mode (d15) define the actuation of the material. Figure 3a shows the
piezoelectric actuation mode with the six orientations of the coordinate systems (x, y, z, θx,
θy, and θz) and the polarization of a single layer piezoelectric crystal under P. For d33 and d31
modes, E applied parallel to P results in a longitudinal deformation (δh) and a transversal
deformation (δl) simultaneously (Figure 3b), whereas E is perpendicular to P for d15 and
produces shear deformation (δs) (Figure 3c) [111]. Among these modes, d33 and d31 provide
higher strains than d15. Piezoelectric actuators produce small strains in an accurate and fast
manner, and thus have been used for a variety of high-precision actuation applications such
as micro/nano-positioning systems [112, 113], micropumps [114, 115], and micro-robotics
[116]. In spite of their advantages, incorporation of piezoelectric materials such as lead
zirconate titanate (PZT) and lead magnesium niobate-PZT ceramics in MEMS fabrication is
often challenging due to the need for high-temperature thermal processes and the instability
of deposited materials [117, 118]. Besides, the need for relatively high driving voltages and
The published version is available at https://doi.org/10.1088/1361-6439/ab8832
8
the large hysteresis/nonlinearity are other factors that can limit their application range [105,
119, 120].
2.4 Fluid-Driven Actuators
Soft and flexible actuators have been attracting attention for biomedical applications as tissue
interaction with mechanically rigid actuators could lead to damage to the tissue. In this
context, many studies have looked at hyperelastic-material-based pneumatic and hydraulic
actuators. These types of actuators are typically comprised of fibreless or fibre-reinforced
polymeric channel structures that allow for supply of gas or liquid (typically air or water,
respectively) to the channels [121] (e.g., McKibben artificial muscle [122]). Once fluidic
pressure is applied to the actuator’s channel, it causes elastic deformation in its overall
structure, resulting in a designed mode of actuation such as expansion, contraction, bending
or twisting motions [123, 124]. For example, pneumatic actuators having symmetric cross
sections expand or contract, while those with asymmetric cross sections (created by, e.g.,
bonding two flexible layers with different wall thicknesses or stiffness levels), such as
Figure 3. Piezoelectric actuation modes:
(a) Orientations of the actuation field and
polarization field; (b) longitudinal and transversal modes; (c) shear mode.
The published version is available at https://doi.org/10.1088/1361-6439/ab8832
9
pneumatic balloon actuators (PBA), show bending deformations [125]. Likewise, two arrays
of PBAs combined in the opposite bending directions cause twisting motions (Figure 4) [126,
127]. In addition, the pneumatic actuator with a single or dual-channel structure can produce
bidirectional curling or bending motions, respectively [128-131], while a three-channelled
pneumatic soft actuator offers bending motions in up to six different directions [132]. Based
on these features, Suzumori et al. developed a flexible microactuator having three chambers
for pneumatic supply [133, 134]. The actuator had a cylindrical fibre-reinforced rubber
structure that provided 3-degree-of-freedom motions. Another study investigated a MEMS-
based hydraulic actuator based on a finger-shaped chamber structure for its actuation, which
used an integrated heater to pressurize the fluid through its thermal expansion [135]. The
fluid-driven actuators offer advantageous features such as high flexibility, large displacement,
biocompatibility (when fabricated/coated with biocompatible materials), lightweight, high
power-to-weight ratio, simple/low-cost fabrication [136, 137], which makes them suitable for
Figure 4. Overview of flexible pneumatic actuators showing four different actuation
modes. Reproduced with permission. [127] Copyright 2014, Elsevier.
The published version is available at https://doi.org/10.1088/1361-6439/ab8832
10
applications in medical and surgical devices, whereas the need for means of fluid supply and
pressurization, actuation precision, and miniaturization are general areas of limitations.
3. Biomedical Applications of MEMS Actuators
This section emphasizes the applications of the aforementioned actuators in biomedical areas
with a focus on lab-on-a-chip, drug delivery systems, cardiac devices, and surgical tools. The
key functions of reported devices and the particular contributions of microactuators to them
are discussed.
3.1. Lab-on-a-Chip (LoC)
LoC is a class of miniaturized microfluidic devices configured in a single-chip form that is
primarily designed for biological or chemical processing and analysis [138]. These devices
allow miniaturization and amalgamation of complex processes to be implemented on a small
chip, which otherwise needs to be operated via repetitive laboratory tasks. The key features of
these devices include compactness/portability, dramatic reduction of required chemicals and
samples, higher process controllability, and faster analysis. The parallelization of many
functions integrated on LoC is leading to an emerging trend in point-of-care diagnostics
[139]. LoC devices are functionalized by forced fluid flow through microfluidic channels
patterned on them. To control flow sequence, duration and timing, direction, and flow rate of
each fluid being processed, micro-scale pumps, and valves are integrated with the channels
on the chip, allowing for precise on-chip manipulation of small quantities of particular fluids.
Piezoelectric actuators have been one type of the actuators widely used as micropump
elements in LoC to control the fluid flow with high accuracy. For example, a multi-chamber
piezoelectric pump was reported to control the fluid flow rate [140] (Figure 5a). As a
sinusoidal signal was applied to the actuator, the chamber expanded and opened the valve,
causing the fluid flow based on the inverse piezoelectric actuation. For point-of-care testing
and chemical analysis, a plug-and-play microfluidic chip integrated a piezoelectric peristaltic
micropump was demonstrated [141]. The fluid in the microchannel was transported through
impacting actions provided by the piezoelectric actuator (Figure 5b). In order to enhance the
functionality and performance of LoCs, researchers have also incorporated surface acoustic
wave (SAW) driven piezoelectric actuators into the LoCs to precisely control fluid flows and
microparticles. SAW based actuators are advantageous in LoCs owing to their features such
as low cost, simple fabrication, fast actuation, high adaptability, contact-free particle
The published version is available at https://doi.org/10.1088/1361-6439/ab8832
11
manipulation, and biocompatibility [142]. For instance, Ding et al. demonstrated standing
SAW based acoustic tweezers to trap and manipulate single microparticles, cells, and
organisms in a microfluidic chip. These tweezers were shown for real-time manipulation of
microparticles by utilizing a wide resonance band of interdigitated transducers [143]. For the
fluid-driven actuation approach, a LoC based on thermo-pneumatic actuation was reported to
control the flow rate inside the microfluidic channel (Figure 5c) [144]. In addition, a multi-
throughput multi-organ-on-a-chip system was developed by utilizing a pneumatic actuator
Figure 5. LoC systems and their components. (a) Schematic and prototype of a piezo-
actuated pump. Reproduced with permission [140].
Copyright 2019, Elsevier. (b)
Piezoelectric-actuator-based microfluidic pump module. Reproduced with permission
[141]. Copyright 2019, Elsevier. (c) Thermopneumatically
actuated microchamber.
Reproduced with permission [144]. Copyright 2019, Elsevier. (d) Pneumatically driven
multi-organ-on-a-plate system, showing (top) culture device, (middle) microfluidic plates,
(bottom left) culture unit and Laplace valves, and (bottom right) membrane insert and
culture chamber. Reproduced with permission [145]. Copyright 2019, RSC publishing.
The published version is available at https://doi.org/10.1088/1361-6439/ab8832
12
(Figure 5d) [145]. This device could handle eight different conventional cell culture
experiments (including cell seeding, medium change, live/dead staining, cell growth analysis,
and gene expression analysis of collected cells) at a time offering a potential for drug
discovery applications.
Electromagnetic actuators are another group that has been employed in micropump
and microfluidic applications exploiting their favorable features for LoC such as rapid
response, large force, and low-voltage operation. For instance, Pradeep et al. developed an
electromagnetically actuated valves to control multiple fluid flow on a programmable
microfluidics platform (Figure 6a) [146]. The device was comprised of polydimethylsiloxane
(PDMS) based microfluidic channels and membranes with an electronic board that held
solenoids. The activation of the solenoid attracted the valve to deflect the PDMS membrane,
which in turn created a path for fluid flow. Another electromagnetically actuated micropump
was reported to provide bidirectional flow [147]. This device used two pairs of power
inductor and NdFeB magnet (Figure 6b), in which the two magnets were synchronously
Figure 6. Electromagnetically actuated microfluidic devices. (a) Schematic and image of
fabricated microfluidic channel with active valves. Reproduced with permission [146].
Copyright 2018, Elsevier. (b) Schematic diagrams of (left) a dual-chamber micropump
and (right) operating principle of the actuation with positive and negative driving
voltages. Reproduced with permission [147]. Copyright 2018, Elsevier.
The published version is available at https://doi.org/10.1088/1361-6439/ab8832
13
actuated under either attractive or repulsive condition (by switching the polarity of voltage
applied to the inductors) to pump the fluid inside the channel in either direction. In another
example, Tahmasebipour et al. fabricated an electromagnetic uni-/bi-directional diffuser
micropump, which used the magnetic membrane based on a PDMS-Fe3O4 nanocomposite for
its electromagnetic actuation to create fluid flow through microchannels [148]. These
micropump devices could be employed in various microfluidic and LoC devices.
3.2. Implantable Drug Delivery Systems
Advances in MEMS and miniaturization technologies have enabled implantable biomedical
devices specifically designed to assist in the diagnosis and treatment of chronic or acute
diseases. Micromachined drug delivery systems are among those emerging implantable
devices. Many of these systems are comprised of micro reservoirs that store liquid-phase
drugs and microactuators that constitute a mechanism to eject the drugs out of the systems
and deliver them to the implanted sites [149]. Aside from the significant improvement in
bioavailability of drugs, the advancement of this type of systems is expected to enable
patient-tailored, pin-point treatments of targeted diseases such as cancer, diabetes, and
osteoporosis, while significantly reducing in-vivo invasiveness of the systems due to their
miniaturized forms.
MEMS drug delivery systems use microvalves to channel/regulate the drug flow into
the diseased location [150]. Thermoresponsive hydrogels have been often used to form smart
microvalves in them [84, 86, 151-156]. A study reported an implantable drug delivery device
that was fabricated to integrate PNIPAM microvalves with a wireless resonant heater and a
drug reservoir [84]. The microvalves were patterned using an in-situ photolithography
technique and were wirelessly operated by activating the resonant heater using a tuned
external radiofrequency (RF) field. This hydrogel microvalve demonstrated 38% shrinkage in
its size upon activation that allowed for release of test drug from the reservoir. Another drug
delivery system using a thermoresponsive hydrogel valve was reported to demonstrate its
repeatable drug release mechanism controlled by induction heating [152]. This device
showed the release of drug as well as its reverse flow to refill the reservoir. A more
comprehensive study on drug delivery through a MEMS device using reversible or
irreversible polymeric valves reported reproducible release control utilizing hydrogel-based
artificial muscle [153]. Eddington et al. developed a drug delivery device by employing an
array of pH-sensitive hydrogels (Figure 7a) [154]. Besides above efforts, various studies have
The published version is available at https://doi.org/10.1088/1361-6439/ab8832
14
reported hydrogel-based microvalves that could be applied to MEMS-based drug delivery
[155-159]. As a different approach, piezoelectric microvalves have also been studied for the
same purpose. This was demonstrated, for example, in a study that developed a wirelessly
controlled normally-closed piezoelectric microvalve activated by an inductor-capacitor (LC)
resonant circuit (Figure 7b) [160]. The activation of the LC circuit required the field
frequency to be modulated to 10 kHz resonant frequency that matched the optimal operating
frequency of the device.
Micropumps are another essential element for MEMS drug delivery systems to
transport drugs from the reservoirs to the outlets of the systems. SMA, thermopneumatic and
piezoelectric actuators have been among those often used in micropump-driven systems. An
implantable drug delivery chip reported in [70] integrated an SMA-based micropump for the
release of stored drug from the chip. The SMA was bulk-micromachined to form a resonant
circuit, which served as a self-heat source activated by RF power transfer to allow frequency-
selective actuation and pumping of drug out of the chip. Thermopneumatic micropumps
based on a similar powering method were developed for release control [161], including
multiple drug delivery and mixing with a zigzag micromixer [162]. Piezoelectric actuated
micropumps were also reported for implantable drug delivery applications [163, 164].
Besides, a polymer-based reusable implantable drug delivery system with refillable
functionality was developed [165]. This device was designed to provide control and refillable
functionalities for broad drug compatibility. Some of the implantable drug delivery systems
were reported to integrate SMP actuators [72, 166, 167]. For example, studies reported the
SMP-pumped implantable device operated by external RF magnetic fields with an actuation
range of 140µm using a 50-mW RF power and showed an average release rate of 0.172
µL/min [72, 166]. A chemotherapy drug release system was realized using hydrolytic
degradable SMPs and was evaluated in the impact of the drug release profile [167]. Apart
from the actuation mechanisms discussed above, electrochemically driven micropumps have
been shown in several reports [168-171]. These studies integrated an electrochemical bellow
actuator, transcutaneous cannula, and a dual regulation valve to form an implantable drug
delivery device [168], showing in-vivo implementation for anti-cancer drug delivery through
wireless powering [169], and demonstrated similar devices for controlled delivery of boluses
from the fabricate prototypes (Figure 7c) [170, 171].
The published version is available at https://doi.org/10.1088/1361-6439/ab8832
15
3.3. Cardiac Devices
Many implantable devices are targeted at providing enhanced diagnoses and/or therapeutic
treatments for specific diseases in vivo. Cardiac implants are a good example of them.
Figure 7. Drug delivery microsystems: (a) (Left) complete microfluidic device and (right)
integrated array of hydrogel actuators. Reproduced with permission [154]. Copyright
2004, IEEE. (b) (Top left) schematic and cross-sectional diagrams of the device, and
fabrication results showing (top right) top and bottom molds and (bottom) device under
off and on states. Reproduced with permission [160]. Copyright 2018, Elsevier. (c) (Left)
schematic diagram and (right) fabrication result of wirelessly powered electrochemical
bellow micropump. Reproduced with permission [171]. Copyright 2016, Elsevier.
The published version is available at https://doi.org/10.1088/1361-6439/ab8832
16
Atherosclerosis is a type of cardiovascular disease where arteries become hardened and
narrowed due to plaque build-up on their inner walls. In conjunction with balloon angioplasty
to treat atherosclerosis, the endovascular mechanical implants called stents are commonly
used as chronic vascular scaffolds to keep the blood vessel open. Most of commercially
available stents are metallic, made of biocompatible alloys such as medical-grade stainless
steel and Nitinol, to configure balloon-expandable or self-expanding stents. These stents with
mesh-like walls are manufactured by laser micromachining of the specific alloy tubes. The
deployment of the self-expanding stents in arteries relies on thermoresponsive actuation of
Nitinol [170]. The stent is positioned at the target location via the delivery catheter and then
(by removing the covering sheath) allowed to self-expand to its memorized diameter through
the martensite-to-austenite phase transformation upon exposure to the body temperature [172,
173]. After their implantation, expanded stents experience elastic recoil of blood vessels,
which can lead to their mechanical failures, a continuing issue for these implants. As an
approach to address this type of failure, a Nitinol-based actuator called the recoil-resilient
ring was investigated to show its ability to improve the radial stiffness of stents when
integrated with them [174]. A newer work demonstrated multiple stage expansion of SMA-
based stent via wireless RF control aiming to address recoil and restenosis issues of stents
[175]. While not as extensive as the case of SMA, the use of SMP has also been investigated
in several studies towards self-expanding stent applications. For example, one study
presented a synthesized SMP for stent application, reporting that the polymer showed 100%
strain recovery [176]. The device displayed high rubbery shear moduli in the range of 2 MPa
and the constrained stress-strain recovery cycle showed very low hysteresis. Another work
presented a biodegradable and self-expandable SMP stent showing excellent mechanical
properties as well as biocompatibility [177].
Thermal therapy commonly known as hyperthermia is a noninvasive technique that
has been used to kill cancerous cells [178]. This therapeutic approach was also reported to be
effective in suppressing the occurrence of restenosis, the most common post-stenting
complication, and following this path, stent-based endohyperthermia was investigated to
enable post-stenting thermal stimulation in a wireless manner [179]. This active “hot” stent
was designed to electrically resonate when exposed to a RF field and implement frequency-
selective heating for vascular treatment. A stent-hyperthermia system based on this principle
was demonstrated through animal tests [180, 181]. To circumvent overheating of the stent
device under excitation, a biocompatible MEMS circuit breaker chip was developed and
integrated with the hot stents (Figure 8a) [182]. This circuit breaker chip functioned as a
The published version is available at https://doi.org/10.1088/1361-6439/ab8832
17
thermoresponsive contact switch with a SMA actuator, or an absolute temperature limiter,
enabling self-regulation of stent’s resonance and thus temperature [182, 183]. Figure 8b
shows an expansion process of the integrated stent device demonstrating automatic switching
and overheat prevention when wirelessly powered [184]. The reported circuit breaker chip
was claimed to be used for temperature regulation of other types of electronic implants.
Aside from stent related applications, shape memory materials have been utilized in
other cardiac devices that exploit their actuation and deployment triggered by the body
temperature. One example is the SMP-based rings that have been used for cardiac valve
repair to reduce mitral regurgitation [185]. Closure devices have been widely used in
intervention treatment for congenital heart disease that is known as abnormal anatomy caused
by dysplasia. Several studies were reported to develop Nitinol-based closure devices (Figure
9) [186]. This type of devices is delivered into the body using its delivery system in a
compressed state and then deployed to its original shape at the target location. A well-known
occlude device was realized with two Nitinol woven discs for closure of congenital heart
defects [187].
3.4 Surgical and Endoscopic Tools
MEMS actuators offer promising opportunities in creating novel surgical devices as well. In
particular, these actuators based on shape-memory materials, piezoelectric, and pneumatic
Figure 8. RF-powered resonant “hot” stent for wireless restenosis treatment. (a) MEMS
circuit-breaker microchip for self-regulation of stent temperature.
Reproduced with
permission [182]. Copyright 2017, IEEE. (b) Deployment of the stent device with circuit-
breaker microchip. Reproduced with permission [184]. Copyright 2015, IEEE.
The published version is available at https://doi.org/10.1088/1361-6439/ab8832
18
principles and related fabrication processes are paving avenues to miniaturizing and
improving the tools for surgical, interventional, and related procedures including catheters,
manipulators, endoscopes, and imaging devices. Utilizing the features of nanometer-range
resolution and fast response, piezoelectric microactuators have been applied for delivering
and scanning high-frequency laser pulses for microsurgery purposes. For example,
Ferhanoglu et al. reported rapid removal of bulk tumors and bones using the 5-mm-diameter
fiber device comprised of an air-core photonic bandgap fiber for delivery of high energy laser
pulses, a piezoelectric tube actuator for fiber scanning, and two aspheric lenses for focusing
the laser beam [188]. To enhance the visualization of fine biopsy needles under ultrasound
imaging, the needle-like catheter that equipped a miniaturized ultrasonic actuator was
developed with a PZT layer sandwiched between two flexible electrodes using MEMS
technology [189]. Being attached to a catheter, the actuator radiated low-intensity ultrasound
for detection of a biopsy needle tip under sonography. Likewise, to perform a non-abdominal
operation or microsurgery, a micro ultrasonic scalpel was developed using PZT deposited
through a hydrothermal method [190, 191].
Pneumatic actuators shaped with soft and flexible elastomers are considered as one of
Figure 9. Nitinol-based closure devices for congenital heart disease: (a) Amplatzer ASD
Occluder; (b) Occlutech Figulla ASD Occluder. Reproduced with permission [186].
Copyright 2019, Elsevier.
The published version is available at https://doi.org/10.1088/1361-6439/ab8832
19
the most suitable candidates for surgical device applications. Many studies utilized the
anisotropic rigidity of PBA in developing bending actuators for active catheter tools. For
example, Ruzzu et al. reported a system for fixing and orientating the catheter tip consisting
of three inflatable microballoons [192]. The microballoons were mounted on the three sides
of the catheter tip and controlled by electro-thermo-pneumatic microvalves. When deflated,
these balloons exerted a force on the wall of the vessel, causing a change in the position and
orientation of the catheter tip. In addition, a telescopic motion was achieved by connecting
several PBA pairs in series in order to actuate commercial forceps [193]. Another PBA-based
device with a cylindrical microstructure was developed to solve a bubbling problem in the
intestinal tract, which caused undesirable stagnation blocking the observation of cells [194].
Supplying air to the artificial intestinal tract via microchannel, the PBA gradually
transformed from flat to circular tube that allowed perfusion of the culture media. For
endoscopic fluorescence imaging and diagnosis, a flexible end-effector was developed via
integration of a PDMS-based PBA, serving as scanning actuator, with an SU-8 optical
waveguide using MEMS fabricated techniques [195].
Besides the PBA-based approaches, various efforts have tailored pneumatic and other
fluid-driven actuators to develop different surgical tools. For example, a pneumatically
actuated micro-gripper was reported to manipulate embryos for cloning applications, (Figure
10a) [196]. The micro-gripper consisted of two main parts; the micro pneumatic chamber
with a flexible membrane and the hinged gripper arms connected to the membrane. Supplying
pressurized air to the membrane, it deflected both the arms to provide a gripping motion.
Traditional laparoscopes used for certain surgical interventions (such as total mesorectal
excision) lack a flexibility sufficient to safely maneuver and reach difficult surgical targets.
This need was approached through the development of the robotic device composed of two
pneumatically actuated identical modules, capable of omnidirectional bending and
elongation, to allow for highly dexterous and safe navigation [197, 198]. Becker et al.
developed a tissue retraction device for treatment of lesions in the gastrointestinal tract [199].
This device was comprised of three main integrated components, i.e., a rigid expandable
geometric structure, inflatable pneumatic actuators, and a vacuum gripper fabricated using the
pop-up book MEMS technique. Similarly, to improve the distal dexterity and enable tissue
retraction, the soft pop-up actuators were exploited to form a multi-articulated robotic arm
(Figure 10b) [200]. Here, the millimeter-scale hybrid soft pop-up actuators were embedded
with capacitive sensing elements to achieve proprioceptive actuation. Endoscopic devices
also often suffer from limited distal tip dexterity, and this issue has been tacked by
The published version is available at https://doi.org/10.1088/1361-6439/ab8832
20
incorporating pneumatic actuation mechanisms with them. For example, pneumatic tubular
actuators were developed and optimized for applications in flexible microactuator-based
endoscopes to facilitate colonoscopy [201, 202] as well as a bronchoscope to observe lung
airway and obstructions in the bronchus [203]. Combining a chip-on-tip CMOS camera with
an elastic inflatable microactuator, Gorissen et al. presented a flexible endoscope for
navigating through intricate topologies of the human body [204].
The endoscopic devices with active scanning functions have been developed by
adopting different actuation methods besides pneumatic one. For example, to obtain in-vivo
local images for tissue diagnostics, an active optical coherence tomography (OCT) probe was
developed with two-axis scanning electrothermal MEMS micromirror, gradient refractive
Figure 10. MEMS-enabled surgical and endoscopic tools. (a) Pneumatically actuated
micro-gripper. Reproduced with permission [196]. Copyright 2015, Elsevier. (b)
Conceptual 3D model and optical image of the soft pop-up actuator. Reproduced with
permission [200]. Copyright 2014, Elsevier. (c) Electrothermally actuated MEMS
scanning mirror for OCT probe and optical images of fabricated micromirrors.
Reproduced with permission [205]. Copyright 2008, IOP Publishing. (d) MEMS-based 3D
confocal scanning microendoscope. Reproduced with permission [207]. Copyright 2013,
Elsevier. (e) Side-viewing Raman probe with integrated MEMS rotary motor. Reproduced
with permission [208]. Copyright 2019, Wiley-VCH.
The published version is available at https://doi.org/10.1088/1361-6439/ab8832
21
index lens, and single-mode fiber integrated on silicon optical bench (SiOB) substrate (Figure
10c) [205]. For three-dimensional (3D) imaging, a 2-axis MEMS mirror with a preset (45°)
angle was directly integrated on a SiOB. The probe was enclosed within a biocompatible,
transparent and waterproof polycarbonate tube for in-vivo applications. A similar active OCT
probe enabled by an electrothermal MEMS mirror was also reported for real-time imaging of
internal organs such as gastrointestinal tract, stomach, small intestine, and esophagus [206].
The unique features of this MEMS-mirror design were a large scan range of ±30°, a high
speed of about 2.5 frames per second, and a body-safe driving voltage of 5.5 V. Following
the same scanning approach, a fiber-optic 3D microendoscope with a confocal scanning
function was developed for early cancer diagnosis (Figure 10d) [207]. The probe was
comprised of electrothermal MEMS scanning mirrors that offered a large imaging field via
both lateral and axial scans with low driving voltages. For endoscopic probes, full
circumferential scanning around the probe is an important ability for screening and detecting
lesions on the walls of luminal organs without blind spots; however, this need is difficult to
meet with 2D MEMS scanners. A tubular MEMS rotary motor was developed for this
application segment and enabled a side-viewing Raman spectroscopy (RS) probe (Figure
10e) [208-210]. This electromagnetic MEMS motor, developed using a self-sustained
ferrofluid bearing in the catheter tube, provided both stepping and continuous rotations of a
probing laser beam and demonstrated full 360° tissue imaging/analysis via RS [208] as well
as OCT [210] modalities ex vivo and in vivo. The motor was also engineered to provide
hydraulic axial motion in addition to rotation for 3D luminal imaging without requiring an
external probe positioning system [209].
4. Conclusion
Continuous advancement of microactuator technologies, along with their fabrications and
integration methods, has led to the emerging areas of biomedical microsystems including
smart implants and surgical devices in miniaturized forms. The success in a targeted
application critically relies on the appropriate selection of a particular actuator, which
depends on various factors besides the fundamental performance of the actuator itself,
including powering and control methods, biocompatibility, level of required packaging, and
cost effectiveness. With an aim to facilitate the development of this emerging field while
addressing those key factors, this paper has presented a comprehensive review of the MEMS
actuators investigated for their biomedical uses with a focus on several common transduction
The published version is available at https://doi.org/10.1088/1361-6439/ab8832
22
types. Table 1 presented a clear comparison of these actuation techniques and their
characteristics. The use of thermoresponsive materials is a promising route to enabling smart
actuation functions with simple designs, an advantageous feature towards device
miniaturization. Among them, SMA offers large displacement and force whereas SMP
possesses relatively high recoverable strain levels. The PNIPAM hydrogel can be compatible
with the standard photo-patterning process and allows for adjustment of its temperature
threshold. The above attributes often make them suitable for applications in drug delivery,
cardiac and surgical devices. The electromagnetic actuators with their large displacement,
fast response and low-voltage powering features are usable for the development of LoC and
their active elements such as micropumps and microvalves. The piezoelectric actuators are a
powerful enabling technology for devices targeted at micro/nano-scale positioning,
micropumps, and micro-robotics. Being softer and flexible, the fluid-driven actuators offer a
variety of application opportunities in surgical devices.
Exploiting these favorable features, thermoresponsive, electromagnetic, and
piezoelectric actuators are widely applied for implantable devices. However, they require an
attention in a few factors. For their medical and implant applications, these actuators are often
powered using batteries. This may cause not only the need for periodic replacement through
surgical procedure but also significantly increase the overall device sizes and hence their
invasiveness in the body. While wireless powering and control methods for smart implants
are being widely explored, the issues around their efficiency, reliability, and
biocompatibility/safety will need to be addressed. The safety factor includes proper heat
management and necessary packaging that, in turn, can negatively impact on the device
performance and size. One of the key approaches to addressing powering issues would be in-
situ energy harvesting from the implanted environment, which may be achieved using similar
principles of some of the abovementioned MEMS actuators but with reversed transductions
converting environmental stimuli to electrical energy.
The published version is available at https://doi.org/10.1088/1361-6439/ab8832
23
Types
Working
principle
Advantages
Disadvantages
Energy
density
(J/m3)
Efficiency
(%)
MEMS Applications
Thermoresponsive
actuators
SMA
Shape-
memory effect
Large displacement
Large force
High mechanical
robustness
Corrosion-resistant
Slow temporal
response
High power
consumption
~107
[211]
<10
[212, 213]
Surgical tools [22, 68]
Implantable devices [65]
Microgrippers [60, 67]
Micropumps [60, 63, 70]
SMP
Shape-
memory effect
Structural flexibility
Large strains
Low density
Tunable transition
temperature
Biodegradable
properties
Slow temporal
response
Low recovery
stress
2-6×105
[214]
<10 [213]
Endovascular devices [74, 75,
176, 177, 185]
Drug delivery devices [74,
166, 167]
Temperature-
Sensitive
Hydrogels
Phase
transition
Tunable
degradation features
Tunable mechanical
features
UV-sensitive
Slow temporal
response
3.5×105
[211]
1.32 [215]
Surgical tools [85]
Microvalves [153-157]
Drug delivery [84, 86, 88-90,
151, 153]
Electromagnetic actuator
Magnetization
effect
Simple drive mode
No nonlinear effect
High field energy
density
Fast response
Large deflection at
low input voltage
High power
dissipation for
driving coils
Volumetric scaling
of produced
electromagnetic
forces that rapidly
drop as the device
size shrinks
4×106
[213]
>90
[212, 213]
Microgrippers [100]
Micropumps [101-103]
The published version is available at https://doi.org/10.1088/1361-6439/ab8832
24
Table 1. Performance comparison of actuation techniques
Parasitic loss at
high frequency
Microvalves [104]
Piezoelectric actuator
Piezoelectric
effect
Fast response
High displacement
resolution
High efficiency
Compact structure
Immunity to
magnetic field
Require high-
temperature thermal
processes for
incorporation of
piezo materials
High driving
voltage
Large hysteresis
nonlinearity
105 [213]
>90
[212, 213]
Micropumps [114, 115]
Micro-robotics [116]
Fluid-Driven actuator
Elastic
deformation
High flexibility
Large displacement
Lightweight
High power-to-
weight ratio
Simple/low-cost
fabrication
Low force exertion
Limited number of
degree of freedom
1.2×106
[216]
30-40
[212, 213]
Medical and surgical devices
[136, 137]
The published version is available at https://doi.org/10.1088/1361-6439/ab8832
25
Acknowledgments
This work is partially supported by Universiti Teknologi Malaysia under UTM Shine
Program (GUP 17H61) and Prototype Development Fund (UTMPR 00L28), and Ministry of
Education Malaysia under Prototype Development Research Scheme (PRGS 4L690) and the
Fundamental Research Grant Scheme (FRGS/1/2019/TK05/UNIM/02/2).
References
[1] A. C. R. Grayson et al., "A BioMEMS review: MEMS technology for physiologically
integrated devices," Proceedings of the IEEE, vol. 92, no. 1, pp. 6-21, 2004.
[2] R. S. Shawgo, A. C. R. Grayson, Y. Li, and M. J. Cima, "BioMEMS for drug delivery,"
Current Opinion in Solid State and Materials Science, vol. 6, no. 4, pp. 329-334, 2002.
[3] E. E. Nuxoll and R. A. Siegel, "BioMEMS devices for drug delivery," IEEE Engineering in
Medicine and Biology Magazine, vol. 28, no. 1, pp. 31-39, 2009.
[4] W. Wang and S. A. Soper, Bio-MEMS: technologies and applications. CRC press, 2006.
[5] A. Haeberlin et al., "The first batteryless, solar-powered cardiac pacemaker," Heart Rhythm,
vol. 12, no. 6, pp. 1317-1323, 2015/06/01/ 2015.
[6] E. Chow, S. Sanghani, and M. Morris, "Wireless MEMS-based implantable medical devices
for cardiology," in Wireless MEMS Networks And Applications: Elsevier, 2017, pp. 77-100.
[7] G. H. Bardy et al., "Amiodarone or an implantable cardioverter–defibrillator for congestive
heart failure," New England Journal of Medicine, vol. 352, no. 3, pp. 225-237, 2005.
[8] V. Kutyifa et al., "Multicenter Automatic Defibrillator Implantation TrialSubcutaneous
Implantable Cardioverter Defibrillator (MADIT S-ICD): Design and clinical protocol,"
American Heart Journal, vol. 189, pp. 158-166, 2017/07/01/ 2017.
[9] D. Figeys and D. Pinto, "Lab-on-a-chip: a revolution in biological and medical sciences," ed:
ACS Publications, 2000.
[10] V. C. Shukla et al., "Lab-on-a-Chip platforms for biophysical studies of cancer with single-
cell resolution," Trends in biotechnology, vol. 36, no. 5, pp. 549-561, 2018.
[11] G.-P. Nikoleli, C. G. Siontorou, D. P. Nikolelis, S. Bratakou, S. Karapetis, and N. Tzamtzis,
"Biosensors Based on Microfluidic Devices Lab-on-a-Chip and Microfluidic Technology," in
Nanotechnology and Biosensors: Elsevier, 2018, pp. 375-394.
[12] L. S. Mattos, D. G. Caldwell, G. Peretti, F. Mora, L. Guastini, and R. Cingolani,
"Microsurgery robots: addressing the needs of high-precision surgical interventions," Swiss
medical weekly, vol. 146, no. 4344, 2016.
[13] J. Troccaz, G. Dagnino, and G.-Z. Yang, "Frontiers of medical robotics: from concept to
systems to clinical translation," Annual review of biomedical engineering, vol. 21, pp. 193-
218, 2019.
[14] A. Acemoglu, N. Deshpande, and L. S. Mattos, "Towards a magnetically-actuated laser
scanner for endoscopic microsurgeries," Journal of Medical Robotics Research, vol. 3, no. 02,
p. 1840004, 2018.
[15] A. Acemoglu, D. Pucci, and L. S. Mattos, "Design and Control of a Magnetic Laser Scanner
for Endoscopic Microsurgeries," IEEE/ASME Transactions on Mechatronics, vol. 24, no. 2,
pp. 527-537, 2019.
[16] C. A. Carabalí, "Robotic Surgery: State of the art and challenges for the future."
[17] A. Gupta and P. Pal, "8 - Micro-electro-mechanical systembased drug delivery devices," in
Bioelectronics and Medical Devices, K. Pal, H.-B. Kraatz, A. Khasnobish, S. Bag, I.
Banerjee, and U. Kuruganti, Eds.: Woodhead Publishing, 2019, pp. 183-210.
[18] H. J. Lee, N. Choi, E.-S. Yoon, and I.-J. Cho, "MEMS devices for drug delivery," Advanced
drug delivery reviews, vol. 128, pp. 132-147, 2018.
[19] H. Kaji, N. Nagai, M. Nishizawa, and T. Abe, "Drug delivery devices for retinal diseases,"
Advanced drug delivery reviews, vol. 128, pp. 148-157, 2018.
The published version is available at https://doi.org/10.1088/1361-6439/ab8832
26
[20] K. S. Yadav, S. Kapse-Mistry, G. Peters, and Y. Mayur, "E-drug delivery: a futuristic
approach," Drug discovery today, 2019.
[21] L. Li and Z. J. Chew, "Microactuators: Design and technology," in Smart Sensors and MEMs:
Elsevier, 2018, pp. 313-354.
[22] M. R. A. Kadir, D. E. O. Dewi, M. N. Jamaludin, M. Nafea, and M. S. M. Ali, "A multi-
segmented shape memory alloy-based actuator system for endoscopic applications," Sensors
and Actuators A: Physical, vol. 296, pp. 92-100, 2019.
[23] R. E. Pawinanto, J. Yunas, A. Alwani, N. Indah, and S. Alva, "Electromagnetic Micro-
Actuator with Silicon Membrane for Fluids Pump in Drug Delivery System," lab-on a chip,
vol. 1, no. 2, p. 3, 2019.
[24] J. Li, H. Huang, and T. Morita, "Stepping piezoelectric actuators with large working stroke
for nano-positioning systems: a review," Sensors and Actuators A: Physical, 2019.
[25] M. Nafea, J. Baliah, and M. S. M. Ali, "Modeling and simulation of a wirelessly-powered
thermopneumatic micropump for drug delivery applications," Indonesian Journal of
Electrical Engineering and Informatics (IJEEI), vol. 7, no. 2, pp. 182-189, 2019.
[26] B. Rouzbeh, G. M. Bone, and G. Ashby, "High-accuracy position control of a rotary
pneumatic actuator," IEEE/ASME Transactions on Mechatronics, vol. 23, no. 6, pp. 2774-
2781, 2018.
[27] W. Zhao, L. Liu, F. Zhang, J. Leng, and Y. Liu, "Shape memory polymers and their
composites in biomedical applications," Materials Science and Engineering: C, 2018.
[28] A. Lendlein and O. E. Gould, "Reprogrammable recovery and actuation behaviour of shape-
memory polymers," Nature Reviews Materials, p. 1, 2019.
[29] R. W. Mailen, C. H. Wagner, R. S. Bang, M. Zikry, M. D. Dickey, and J. Genzer, "Thermo-
mechanical transformation of shape memory polymers from initially flat discs to bowls and
saddles," Smart Materials and Structures, vol. 28, no. 4, p. 045011, 2019.
[30] M.-T. Ke, J.-H. Zhong, and C.-Y. Lee, "Electromagnetically-actuated reciprocating pump for
high-flow-rate microfluidic applications," Sensors, vol. 12, no. 10, pp. 13075-13087, 2012.
[31] C.-Y. Lee, Z.-H. Chen, H.-T. Chang, C.-Y. Wen, and C.-H. Cheng, "Design and fabrication
of novel micro electromagnetic actuator," Microsystem technologies, vol. 15, no. 8, pp. 1171-
1177, 2009.
[32] K. H. Kim, H. J. Yoon, O. C. Jeong, and S. S. Yang, "Fabrication and test of a micro
electromagnetic actuator," Sensors and Actuators A: Physical, vol. 117, no. 1, pp. 8-16, 2005.
[33] Y. A. Yildirim, A. Toprak, and O. Tigli, "Piezoelectric membrane actuators for micropump
applications using PVDF-TrFE," Journal of Microelectromechanical Systems, vol. 27, no. 1,
pp. 86-94, 2017.
[34] L. Wang, W. Chen, J. Liu, J. Deng, and Y. Liu, "A review of recent studies on non-resonant
piezoelectric actuators," Mechanical Systems and Signal Processing, vol. 133, p. 106254,
2019.
[35] K. R. Oldham, J. S. Pulskamp, R. G. Polcawich, and M. Dubey, "Thin-film PZT lateral
actuators with extended stroke," Journal of Microelectromechanical Systems, vol. 17, no. 4,
pp. 890-899, 2008.
[36] I. A. Ivan, M. Rakotondrabe, P. Lutz, and N. Chaillet, "Quasistatic displacement self-sensing
method for cantilevered piezoelectric actuators," Review of Scientific instruments, vol. 80, no.
6, p. 065102, 2009.
[37] D. Saravanakumar, B. Mohan, and T. Muthuramalingam, "A review on recent research trends
in servo pneumatic positioning systems," Precision Engineering, vol. 49, pp. 481-492,
2017/07/01/ 2017.
[38] H. I. Ali, S. Noor, S. Bashi, and M. Marhaban, "A review of pneumatic actuators (modeling
and control)," Australian Journal of Basic and Applied Sciences, vol. 3, no. 2, pp. 440-454,
2009.
[39] G. K. Klute, J. M. Czerniecki, and B. Hannaford, "McKibben artificial muscles: pneumatic
actuators with biomechanical intelligence," in 1999 IEEE/ASME International Conference on
Advanced Intelligent Mechatronics (Cat. No. 99TH8399), 1999, pp. 221-226: IEEE.
The published version is available at https://doi.org/10.1088/1361-6439/ab8832
27
[40] R. V. Martinez, C. R. Fish, X. Chen, and G. M. Whitesides, "Elastomeric origami:
programmable paper‐elastomer composites as pneumatic actuators," Advanced functional
materials, vol. 22, no. 7, pp. 1376-1384, 2012.
[41] P. S. Gandhi, A. Savalia, and H. Shah, "Design, fabrication, and characterization of a
pneumatic micro actuator," in ASME 2011 International Mechanical Engineering Congress
and Exposition, 2011, pp. 455-461: American Society of Mechanical Engineers Digital
Collection.
[42] F. Pilate, A. Toncheva, P. Dubois, and J.-M. Raquez, "Shape-memory polymers for multiple
applications in the materials world," European Polymer Journal, vol. 80, pp. 268-294, 2016.
[43] Y.-N. Wang and L.-M. Fu, "Micropumps and biomedical applications–A review,"
Microelectronic Engineering, vol. 195, pp. 121-138, 2018.
[44] D. Han et al., "Soft robotic manipulation and locomotion with a 3d printed electroactive
hydrogel," ACS applied materials & interfaces, vol. 10, no. 21, pp. 17512-17518, 2018.
[45] L. Zhang, I. Desta, and P. Naumov, "Synergistic action of thermoresponsive and
hygroresponsive elements elicits rapid and directional response of a bilayer actuator,"
Chemical Communications, vol. 52, no. 35, pp. 5920-5923, 2016.
[46] C. Ma et al., "A multiresponsive anisotropic hydrogel with macroscopic 3D complex
deformations," Advanced Functional Materials, vol. 26, no. 47, pp. 8670-8676, 2016.
[47] K. Iwaso, Y. Takashima, and A. Harada, "Fast response dry-type artificial molecular muscles
with [c2] daisy chains," Nature chemistry, vol. 8, no. 6, p. 625, 2016.
[48] M.-W. Han et al., "Woven type smart soft composite for soft morphing car spoiler,"
Composites Part B: Engineering, vol. 86, pp. 285-298, 2016.
[49] J.-L. Seguin, M. Bendahan, A. Isalgue, V. Esteve-Cano, H. Carchano, and V. Torra, "Low
temperature crystallised Ti-rich NiTi shape memory alloy films for microactuators," Sensors
and Actuators A: Physical, vol. 74, no. 1-3, pp. 65-69, 1999.
[50] J. J. Gill, K. Ho, and G. P. Carman, "Three-dimensional thin-film shape memory alloy
microactuator with two-way effect," Journal of Microelectromechanical Systems, vol. 11, no.
1, pp. 68-77, 2002.
[51] N. Frantz et al., "Shape memory thin films with transition above room temperature from Ni-
rich NiTi films," Sensors and Actuators A: Physical, vol. 99, no. 1-2, pp. 59-63, 2002.
[52] E. Makino, T. Mitsuya, and T. Shibata, "Fabrication of TiNi shape memory micropump,"
Sensors and Actuators A: Physical, vol. 88, no. 3, pp. 256-262, 2001.
[53] Q. Pan and C. Cho, "The investigation of a shape memory alloy micro-damper for MEMS
applications," Sensors, vol. 7, no. 9, pp. 1887-1900, 2007.
[54] J. K. Paik, E. Hawkes, and R. J. Wood, "A novel low-profile shape memory alloy torsional
actuator," Smart Materials and Structures, vol. 19, no. 12, p. 125014, 2010.
[55] S. Braun, N. Sandstrom, G. Stemme, and W. van der Wijngaart, "Wafer-scale manufacturing
of bulk shape-memory-alloy microactuators based on adhesive bonding of titaniumnickel
sheets to structured silicon wafers," Journal of microelectromechanical systems, vol. 18, no.
6, pp. 1309-1317, 2009.
[56] A. AbuZaiter, M. Nafea, A. A. M. Faudzi, S. Kazi, and M. S. M. Ali, "Thermomechanical
behavior of bulk NiTi shape-memory-alloy microactuators based on bimorph actuation,"
Microsystem Technologies, vol. 22, no. 8, pp. 2125-2131, 2016.
[57] S. Jayachandran et al., "Investigations on performance viability of NiTi, NiTiCu, CuAlNi and
CuAlNiMn shape memory alloy/Kapton composite thin film for actuator application,"
Composites Part B: Engineering, vol. 176, p. 107182, 2019.
[58] N. Choudhary and D. Kaur, "Shape memory alloy thin films and heterostructures for MEMS
applications: A review," Sensors and Actuators A: Physical, vol. 242, pp. 162-181, 2016.
[59] M. S. M. Ali, B. Bycraft, A. Bsoul, and K. Takahata, "Radio-controlled microactuator based
on shape-memory-alloy spiral-coil inductor," Journal of microelectromechanical systems,
vol. 22, no. 2, pp. 331-338, 2012.
[60] M. A. Zainal, S. Sahlan, and M. S. M. Ali, "Micromachined shape-memory-alloy
microactuators and their application in biomedical devices," Micromachines, vol. 6, no. 7, pp.
879-901, 2015.
The published version is available at https://doi.org/10.1088/1361-6439/ab8832
28
[61] A. AbuZaiter and M. S. M. Ali, "Analysis of thermomechanical behavior of shape-memory-
alloy bimorph microactuator," in 2014 5th International Conference on Intelligent Systems,
Modelling and Simulation, 2014, pp. 390-393: IEEE.
[62] A. AbuZaiter, E. L. Ng, M. S. M. Ali, and S. Kazi, "Miniature parallel manipulator using
TiNiCu shape-memory-alloy microactuators," in 2015 10th Asian Control Conference
(ASCC), 2015, pp. 1-4: IEEE.
[63] F. Sassa, Y. Al-Zain, T. Ginoza, S. Miyazaki, and H. Suzuki, "Miniaturized shape memory
alloy pumps for stepping microfluidic transport," Sensors and Actuators B: Chemical, vol.
165, no. 1, pp. 157-163, 2012.
[64] H. Kato and K. Sasaki, "Transformation-induced plasticity as the origin of serrated flow in an
NiTi shape memory alloy," International Journal of Plasticity, vol. 50, pp. 37-48, 2013.
[65] M. H. Elahinia, M. Hashemi, M. Tabesh, and S. B. Bhaduri, "Manufacturing and processing
of NiTi implants: a review," Progress in materials science, vol. 57, no. 5, pp. 911-946, 2012.
[66] M. Dahmardeh et al., "High‐power MEMS switch enabled by carbon‐nanotube contact and
shape‐memory‐alloy actuator," physica status solidi (a), vol. 210, no. 4, pp. 631-638, 2013.
[67] M. M. Ali and K. Takahata, "Frequency-controlled wireless shape-memory-alloy
microactuators integrated using an electroplating bonding process," Sensors and Actuators A:
Physical, vol. 163, no. 1, pp. 363-372, 2010.
[68] M. Es-Souni, M. Es-Souni, and H. Fischer-Brandies, "Assessing the biocompatibility of NiTi
shape memory alloys used for medical applications," Analytical and bioanalytical chemistry,
vol. 381, no. 3, pp. 557-567, 2005.
[69] M. M. Ali and K. Takahata, "Selective RF wireless control of integrated bulk-micromachined
shape-memory-alloy actuators and its microfluidic application," in 2011 IEEE 24th
International Conference on Micro Electro Mechanical Systems, 2011, pp. 1269-1272: IEEE.
[70] J. Fong, Z. Xiao, and K. Takahata, "Wireless implantable chip with integrated nitinol-based
pump for radio-controlled local drug delivery," Lab on a Chip, vol. 15, no. 4, pp. 1050-1058,
2015.
[71] A. AbuZaiter, O. F. Hikmat, M. Nafea, and M. S. M. Ali, "Design and fabrication of a novel
XYθz monolithic micro-positioning stage driven by NiTi shape-memory-alloy actuators,"
Smart Materials and Structures, vol. 25, no. 10, p. 105004, 2016.
[72] M. Zainal, A. Ahmad, and M. M. Ali, "Frequency-controlled wireless shape memory polymer
microactuator for drug delivery application," Biomedical microdevices, vol. 19, no. 1, p. 8,
2017.
[73] C. Liu, H. Qin, and P. Mather, "Review of progress in shape-memory polymers," Journal of
materials chemistry, vol. 17, no. 16, pp. 1543-1558, 2007.
[74] H. Wache, D. Tartakowska, A. Hentrich, and M. Wagner, "Development of a polymer stent
with shape memory effect as a drug delivery system," Journal of Materials Science:
Materials in Medicine, vol. 14, no. 2, pp. 109-112, 2003.
[75] S. H. Ajili, N. G. Ebrahimi, and M. Soleimani, "Polyurethane/polycaprolactane blend with
shape memory effect as a proposed material for cardiovascular implants," Acta biomaterialia,
vol. 5, no. 5, pp. 1519-1530, 2009.
[76] S. M. Hasan, L. D. Nash, and D. J. Maitland, "Porous shape memory polymers: Design and
applications," Journal of Polymer Science Part B: Polymer Physics, vol. 54, no. 14, pp. 1300-
1318, 2016.
[77] E. Radvar and H. S. Azevedo, "Supramolecular peptide/polymer hybrid hydrogels for
biomedical applications," Macromolecular bioscience, vol. 19, no. 1, p. 1800221, 2019.
[78] P. de Almeida et al., "Cytoskeletal stiffening in synthetic hydrogels," Nature
communications, vol. 10, no. 1, p. 609, 2019.
[79] M. Hippler et al., "Controlling the shape of 3D microstructures by temperature and light,"
Nature communications, vol. 10, no. 1, p. 232, 2019.
[80] H. F. Darge, A. T. Andrgie, H.-C. Tsai, and J.-Y. Lai, "Polysaccharide and polypeptide based
injectable thermo-sensitive hydrogels for local biomedical applications," International
journal of biological macromolecules, 2019.
The published version is available at https://doi.org/10.1088/1361-6439/ab8832
29
[81] R. Greiner, M. Allerdissen, A. Voigt, and A. Richter, "Fluidic microchemomechanical
integrated circuits processing chemical information," Lab on a Chip, vol. 12, no. 23, pp.
5034-5044, 2012.
[82] Y. S. Zhang and A. Khademhosseini, "Advances in engineering hydrogels," Science, vol. 356,
no. 6337, p. eaaf3627, 2017.
[83] Y. Yi, A. Zaher, O. Yassine, U. Buttner, J. Kosel, and I. G. Foulds, "Electromagnetically
powered electrolytic pump and thermo-responsive valve for drug delivery," in 10th IEEE
International Conference on Nano/Micro Engineered and Molecular Systems, 2015, pp. 5-8:
IEEE.
[84] S. Rahimi, E. H. Sarraf, G. K. Wong, and K. Takahata, "Implantable drug delivery device
using frequency-controlled wireless hydrogel microvalves," Biomedical microdevices, vol.
13, no. 2, pp. 267-277, 2011.
[85] M. Selvaraj and K. Takahata, "A steerable smart catheter tip realized by flexible hydrogel
actuator," in 2016 IEEE 29th International Conference on Micro Electro Mechanical Systems
(MEMS), 2016, pp. 161-164: IEEE.
[86] E. Sarraf, G. Wong, and K. Takahata, "Frequency-selectable wireless actuation of hydrogel
using micromachined resonant heaters toward implantable drug delivery applications," in
TRANSDUCERS 2009-2009 International Solid-State Sensors, Actuators and Microsystems
Conference, 2009, pp. 1525-1528: IEEE.
[87] V. Sridhar and K. Takahata, "A hydrogel-based wireless sensor using micromachined variable
inductors with folded flex-circuit structures for biomedical applications," in 2008 IEEE 21st
International Conference on Micro Electro Mechanical Systems, 2008, pp. 70-73: IEEE.
[88] D. Schmaljohann, "Thermo-and pH-responsive polymers in drug delivery," Advanced drug
delivery reviews, vol. 58, no. 15, pp. 1655-1670, 2006.
[89] X. J. Loh, P. Peh, S. Liao, C. Sng, and J. Li, "Controlled drug release from biodegradable
thermoresponsive physical hydrogel nanofibers," Journal of Controlled Release, vol. 143, no.
2, pp. 175-182, 2010.
[90] X. Lin, D. Tang, W. Cui, and Y. Cheng, "Controllable drug release of electrospun
thermoresponsive poly (N‐isopropylacrylamide)/poly (2‐acrylamido‐2‐methylpropanesulfonic
acid) nanofibers," Journal of Biomedical Materials Research Part A, vol. 100, no. 7, pp.
1839-1845, 2012.
[91] S. Yang and Q. Xu, "A review on actuation and sensing techniques for MEMS-based
microgrippers," Journal of Micro-Bio Robotics, vol. 13, no. 1-4, pp. 1-14, 2017.
[92] M. Baù, V. Ferrari, D. Marioli, E. Sardini, M. Serpelloni, and A. Taroni, "Contactless
electromagnetic excitation of resonant sensors made of conductive miniaturized structures,"
Sensors and Actuators A: Physical, vol. 148, no. 1, pp. 44-50, 2008.
[93] M. Bau, V. Ferrari, and D. Maroli, "Contactless excitation of MEMS resonant sensors by
electromagnetic driving," in Proceedings of the COMSOL Conference, 2009.
[94] O. Gomis-Bellmunt and L. F. Campanile, Design rules for actuators in active mechanical
systems. Springer Science & Business Media, 2009.
[95] X. Lv, W. Wei, X. Mao, Y. Chen, J. Yang, and F. Yang, "A novel MEMS electromagnetic
actuator with large displacement," Sensors and Actuators A: Physical, vol. 221, pp. 22-28,
2015.
[96] B. Sheeparamatti and V. V. Naik, "Exploration of micro cantilever based electromagnetic
actuator," in 2016 International Conference on Electrical, Electronics, Communication,
Computer and Optimization Techniques (ICEECCOT), 2016, pp. 350-354: IEEE.
[97] M. U. Khan, C. Prelle, F. Lamarque, and S. Büttgenbach, "Design and assessment of a
micropositioning system driven by electromagnetic actuators," IEEE/ASME Transactions on
Mechatronics, vol. 22, no. 1, pp. 551-560, 2016.
[98] V. F.-G. Tseng, J. Li, X. Zhang, J. Ding, Q. Chen, and H. Xie, "An electromagnetically
actuated micromirror with precise angle control for harsh environment optical switching
applications," Sensors and Actuators A: Physical, vol. 206, pp. 1-9, 2014.
[99] C.-H. Ou, Y.-C. Lin, Y. Keikoin, T. Ono, M. Esashi, and Y.-C. Tsai, "Two-dimensional
MEMS Fe-based metallic glass micromirror driven by an electromagnetic actuator," Japanese
Journal of Applied Physics, vol. 58, no. SD, p. SDDL01, 2019.
The published version is available at https://doi.org/10.1088/1361-6439/ab8832
30
[100] D.-H. Kim, M. G. Lee, B. Kim, and Y. Sun, "A superelastic alloy microgripper with
embedded electromagnetic actuators and piezoelectric force sensors: a numerical and
experimental study," Smart materials and structures, vol. 14, no. 6, p. 1265, 2005.
[101] J. Coppeta et al., "A portable and reconfigurable multi-organ platform for drug development
with onboard microfluidic flow control," Lab on a Chip, vol. 17, no. 1, pp. 134-144, 2017.
[102] J. Getpreecharsawas, I. Puchades, B. Hournbuckle, L. Fuller, R. Pearson, and S. Lyshevski,
"An electromagnetic MEMS actuator for micropumps," in Proceedings of the 2nd
International Conference on Perspective Technologies and Methods in MEMS Design, 2006,
pp. 11-14: IEEE.
[103] I. Amrani, A. Cheriet, and M. Feliachi, "Design and experimental investigation of a bi-
directional valveless electromagnetic micro-pump," Sensors and Actuators A: Physical, vol.
272, pp. 310-317, 2018/04/01/ 2018.
[104] M. Capanu, J. G. Boyd, and P. J. Hesketh, "Design, fabrication, and testing of a bistable
electromagnetically actuated microvalve," Journal of Microelectromechanical Systems, vol.
9, no. 2, pp. 181-189, 2000.
[105] F. Ceyssens, S. Sadeghpour, F. Hiroyuki, and R. Puers, "Actuators: Accomplishments,
opportunities and challenges," Sensors and Actuators A: Physical, vol. 295, pp. 604-611,
2019.
[106] H. Shi, J. Chen, G. Liu, W. Xiao, and S. Dong, "A piezoelectric pseudo-bimorph actuator,"
Applied Physics Letters, vol. 102, no. 24, p. 242904, 2013.
[107] Z. Ding, J. Dong, H. Yin, Z. Wang, X. Zhou, and Z. Xu, "Design and experimental
performances of a piezoelectric stick-slip actuator for rotary motion," in IOP Conference
Series: Materials Science and Engineering, 2019, vol. 563, no. 4, p. 042068: IOP Publishing.
[108] R. K. Jain, S. Majumder, and B. Ghosh, "Design and analysis of piezoelectric actuator for
micro gripper," International Journal of Mechanics and Materials in Design, vol. 11, no. 3,
pp. 253-276, 2015.
[109] J. Ma, Y. Hu, B. Li, Z. Feng, and J. Chu, "Influence of secondary converse piezoelectric
effect on deflection of fully covered PZT actuators," Sensors and Actuators A: Physical, vol.
175, pp. 132-138, 2012/03/01/ 2012.
[110] R. S. Fearing, "Micro-actuators for micro-robots: Electric and magnetic," in Handbook of
Sensors and Actuators, vol. 6: Elsevier, 1998, pp. 161-179.
[111] J. Peng and X. Chen, "A survey of modeling and control of piezoelectric actuators," Modern
Mechanical Engineering, vol. 3, no. 01, p. 1, 2013.
[112] X. Chen, C. Su, Z. Li, and F. Yang, "Design of Implementable Adaptive Control for
Micro/Nano Positioning System Driven by Piezoelectric Actuator," IEEE Transactions on
Industrial Electronics, vol. 63, no. 10, pp. 6471-6481, 2016.
[113] O. Aljanaideh, M. Al Janaideh, and M. Rakotondrabe, "Inversion-free feedforward dynamic
compensation of hysteresis nonlinearities in piezoelectric micro/nano-positioning actuators,"
in 2015 IEEE International Conference on Robotics and Automation (ICRA), 2015, pp. 2673-
2678: IEEE.
[114] H.-K. Ma, W.-F. Luo, and J.-Y. Lin, "Development of a piezoelectric micropump with novel
separable design for medical applications," Sensors and Actuators A: Physical, vol. 236, pp.
57-66, 2015.
[115] H. Ma, R. Chen, and Y. Hsu, "Development of a piezoelectric-driven miniature pump for
biomedical applications," Sensors and Actuators A: Physical, vol. 234, pp. 23-33, 2015.
[116] S. Chopra and N. Gravish, "Piezoelectric actuators with on-board sensing for micro-robotic
applications," Smart Materials and Structures, vol. 28, no. 11, p. 115036, 2019.
[117] H. K. Kommepalli, G. Y. Han, C. L. Muhlstein, S. Trolier-McKinstry, C. D. Rahn, and S. A.
Tadigadapa, "Design, fabrication, and performance of a piezoelectric uniflex microactuator,"
Journal of microelectromechanical systems, vol. 18, no. 3, pp. 616-625, 2009.
[118] Y. Jing and J. Luo, "Structure and electrical properties of PMN-PZT micro-actuator deposited
by tape-casting process," Journal of Materials Science: Materials in Electronics, vol. 16, no.
5, pp. 287-294, 2005.
The published version is available at https://doi.org/10.1088/1361-6439/ab8832
31
[119] S. A. Rios and A. J. Fleming, "Design of a charge drive for reducing hysteresis in a
piezoelectric bimorph actuator," IEEE/ASME Transactions on Mechatronics, vol. 21, no. 1,
pp. 51-54, 2015.
[120] Y. Jian, D. Huang, J. Liu, and D. Min, "High-precision tracking of piezoelectric actuator
using iterative learning control and direct inverse compensation of hysteresis," IEEE
Transactions on Industrial Electronics, vol. 66, no. 1, pp. 368-377, 2018.
[121] A. Faudzi, M. Razif, I. Nordin, K. Suzumori, S. Wakimoto, and D. Hirooka, "Development of
Bending Soft Actuator with Different Braided Angles. IEEE," in ASME International
Conference on Advanced Intelligent Mechatronics. Koahsiung, Taiwan, 2012, pp. 11-14.
[122] M. De Volder, A. Moers, and D. Reynaerts, "Fabrication and control of miniature McKibben
actuators," Sensors and Actuators A: Physical, vol. 166, no. 1, pp. 111-116, 2011.
[123] G. Agarwal, N. Besuchet, B. Audergon, and J. Paik, "Stretchable materials for robust soft
actuators towards assistive wearable devices," Scientific reports, vol. 6, p. 34224, 2016.
[124] M. De Volder and D. Reynaerts, "Pneumatic and hydraulic microactuators: a review," Journal
of Micromechanics and microengineering, vol. 20, no. 4, p. 043001, 2010.
[125] S. Konishi, F. Kawai, and P. Cusin, "Thin flexible end-effector using pneumatic balloon
actuator," Sensors and Actuators A: Physical, vol. 89, no. 1-2, pp. 28-35, 2001.
[126] K. Ogura, S. Wakimoto, K. Suzumori, and Y. Nishioka, "Micro pneumatic curling actuator-
Nematode actuator," in 2008 IEEE International Conference on Robotics and Biomimetics,
2009, pp. 462-467: IEEE.
[127] B. Gorissen, T. Chishiro, S. Shimomura, D. Reynaerts, M. De Volder, and S. Konishi,
"Flexible pneumatic twisting actuators and their application to tilting micromirrors," Sensors
and Actuators A: Physical, vol. 216, pp. 426-431, 2014.
[128] M. Deng, A. Wang, S. Wakimoto, and T. Kawashima, "Characteristic analysis and modeling
of a miniature pneumatic curling rubber actuator," International Conference on Advanced
Mechatronic Systems (ICAMechS), vol. 10, pp. 534-539, 2011.
[129] K. Suzumori, S. Endo, T. Kanda, N. Kato, and H. Suzuki, "A Bending Pneumatic Rubber
Actuator Realizing Soft-bodied Manta Swimming Robot," in Robotics and Automation, 2007
IEEE International Conference on, ed, 2007, pp. 4975-4980.
[130] M. R. M. Razif et al., "Two chambers soft actuator realizing robotic gymnotiform swimmers
fin," in Robotics and Biomimetics (ROBIO), 2014 IEEE International Conference on, ed,
2014, pp. 15-20.
[131] T. Rehman, M. Nafea, T. Saleh, and M. S. M. Ali, "PDMS-based dual-channel pneumatic
micro-actuator," Smart Materials and Structures, 2019.
[132] S. Wakimoto, K. Ogura, K. Suzumori, and Y. Nishioka, "Miniature soft hand with curling
rubber pneumatic actuators," in Robotics and Automation, 2009. ICRA '09. IEEE
International Conference on, ed, 2009, pp. 556-561.
[133] K. Suzumori, "Elastic materials producing compliant robots," Robotics and Autonomous
Systems, vol. 18, pp. 135-140, 1996.
[134] K. Suzumori, S. Iikura, and H. Tanaka, "Development of flexible microactuator and its
applications to robotic mechanisms," ed, 2002, pp. 1622-1627.
[135] S. Mutzenich, T. Vinay, and G. Rosengarten, "Analysis of a novel micro-hydraulic actuation
for MEMS," Sensors and Actuators A: Physical, vol. 116, no. 3, pp. 525-529, 2004.
[136] D. Rus and M. T. Tolley, "Design, fabrication and control of soft robots," Nature, vol. 521,
no. 7553, pp. 467-475, 2015.
[137] D. Trivedi, C. D. Rahn, W. M. Kier, and I. D. Walker, "Soft robotics: Biological inspiration,
state of the art, and future research," Applied bionics and biomechanics, vol. 5, no. 3, pp. 99-
117, 2008.
[138] K. Oh, "Lab-on-chip (LOC) devices and microfluidics for biomedical applications," in MEMS
for Biomedical Applications: Elsevier, 2012, pp. 150-171.
[139] Y. Lim, A. Kouzani, and W. Duan, "Lab-on-a-chip: a component view," Microsystem
Technologies, vol. 16, no. 12, pp. 1995-2015, 2010.
[140] T. Peng et al., "A high-flow, self-filling piezoelectric pump driven by hybrid connected
multiple chambers with umbrella-shaped valves," Sensors and Actuators B: Chemical, vol.
301, p. 126961, 2019.
The published version is available at https://doi.org/10.1088/1361-6439/ab8832
32
[141] T. Ma, S. Sun, B. Li, and J. Chu, "Piezoelectric peristaltic micropump integrated on a
microfluidic chip," Sensors and Actuators A: Physical, vol. 292, pp. 90-96, 2019.
[142] X. Ding et al., "Surface acoustic wave microfluidics," Lab on a Chip, vol. 13, no. 18, pp.
3626-3649, 2013.
[143] X. Ding et al., "On-chip manipulation of single microparticles, cells, and organisms using
surface acoustic waves," Proceedings of the National Academy of Sciences, vol. 109, no. 28,
pp. 11105-11109, 2012.
[144] F. Perdigones, M. Cabello, and J. M. Quero, "Single-use impulsion system for displacement
of liquids on thermoplastic-based lab on chip," Sensors and Actuators A: Physical, vol. 298,
p. 111568, 2019/10/15/ 2019.
[145] T. Satoh et al., "A multi-throughput multi-organ-on-a-chip system on a plate formatted
pneumatic pressure-driven medium circulation platform," Lab on a Chip, vol. 18, no. 1, pp.
115-125, 2018.
[146] A. Pradeep, J. Stanley, B. G. Nair, and T. S. Babu, "Automated and programmable
electromagnetically actuated valves for microfluidic applications," Sensors and Actuators A:
Physical, vol. 283, pp. 79-86, 2018.
[147] M. Rusli, P. S. Chee, R. Arsat, K. X. Lau, and P. L. Leow, "Electromagnetic actuation dual-
chamber bidirectional flow micropump," Sensors and Actuators A: Physical, vol. 282, pp. 17-
27, 2018.
[148] M. Tahmasebipour and A. A. Paknahad, "Unidirectional and bidirectional valveless
electromagnetic micropump with PDMS-Fe 3 O 4 nanocomposite magnetic membrane,"
Journal of Micromechanics and Microengineering, 2019.
[149] M. V. Ramesh, K. S. Mohan, and D. Nadarajan, "An in-body wireless communication system
for targeted drug delivery: Design and simulation," in 2014 International Symposium on
Technology Management and Emerging Technologies, 2014, pp. 56-61: IEEE.
[150] A. Gupta and P. Pal, "Micro-electro-mechanical systembased drug delivery devices," in
Bioelectronics and Medical Devices: Elsevier, 2019, pp. 183-210.
[151] S. Rahimi and K. Takahata, "A wireless implantable drug delivery device with hydrogel
microvalves controlled by field-frequency tuning," in 2011 IEEE 24th International
Conference on Micro Electro Mechanical Systems, 2011, pp. 1019-1022: IEEE.
[152] Y. Yi, R. Huang, and C. Li, "Flexible substrate-based thermo-responsive valve applied in
electromagnetically powered drug delivery system," Journal of materials science, vol. 54, no.
4, pp. 3392-3402, 2019.
[153] L.-M. Low, S. Seetharaman, K.-Q. He, and M. J. Madou, "Microactuators toward
microvalves for responsive controlled drug delivery," Sensors and Actuators B: Chemical,
vol. 67, no. 1-2, pp. 149-160, 2000.
[154] D. T. Eddington and D. J. Beebe, "A valved responsive hydrogel microdispensing device with
integrated pressure source," Journal of Microelectromechanical Systems, vol. 13, no. 4, pp.
586-593, 2004.
[155] D. J. Beebe et al., "Functional hydrogel structures for autonomous flow control inside
microfluidic channels," Nature, vol. 404, no. 6778, p. 588, 2000.
[156] A. Baldi, Y. Gu, P. E. Loftness, R. A. Siegel, and B. Ziaie, "A hydrogel-actuated
environmentally sensitive microvalve for active flow control," Journal of
Microelectromechanical Systems, vol. 12, no. 5, pp. 613-621, 2003.
[157] R. H. Liu, Q. Yu, and D. J. Beebe, "Fabrication and characterization of hydrogel-based
microvalves," Journal of microelectromechanical systems, vol. 11, no. 1, pp. 45-53, 2002.
[158] D. Kim and D. J. Beebe, "A bi-polymer micro one-way valve," Sensors and Actuators A:
Physical, vol. 136, no. 1, pp. 426-433, 2007.
[159] Q. Luo, S. Mutlu, Y. B. Gianchandani, F. Svec, and J. M. Fréchet, "Monolithic valves for
microfluidic chips based on thermoresponsive polymer gels," Electrophoresis, vol. 24, no. 21,
pp. 3694-3702, 2003.
[160] M. Nafea, A. Nawabjan, and M. S. M. Ali, "A wirelessly-controlled piezoelectric microvalve
for regulated drug delivery," Sensors and Actuators A: Physical, vol. 279, pp. 191-203, 2018.
The published version is available at https://doi.org/10.1088/1361-6439/ab8832
33
[161] P. S. Chee, M. N. Minjal, P. L. Leow, and M. S. M. Ali, "Wireless powered thermo-
pneumatic micropump using frequency-controlled heater," Sensors and Actuators A:
Physical, vol. 233, pp. 1-8, 2015.
[162] M. Nafea, A. AbuZaiter, S. Kazi, and M. S. M. Ali, "Frequency-controlled wireless passive
thermopneumatic micromixer," Journal of Microelectromechanical Systems, vol. 26, no. 3,
pp. 691-703, 2017.
[163] J. G. Smits, "Piezoelectric micropump with microvalves," in Proceedings., Eighth
University/Government/Industry Microelectronics Symposium, 1989, pp. 92-94: IEEE.
[164] D. Maillefer, H. van Lintel, G. Rey-Mermet, and R. Hirschi, "A high-performance silicon
micropump for an implantable drug delivery system," in Technical Digest. IEEE
International MEMS 99 Conference. Twelfth IEEE International Conference on Micro
Electro Mechanical Systems (Cat. No. 99CH36291), 1999, pp. 541-546: IEEE.
[165] E. Meng, P.-Y. Li, R. Lo, R. Sheybani, and C. Gutierrez, "Implantable MEMS drug delivery
pumps for small animal research," in 2009 Annual International Conference of the IEEE
Engineering in Medicine and Biology Society, 2009, pp. 6696-6698: IEEE.
[166] M. Zainal and M. M. Ali, "Wireless shape memory polymer microactuator for implantable
drug delivery aplication," in 2016 IEEE EMBS Conference on Biomedical Engineering and
Sciences (IECBES), 2016, pp. 76-79: IEEE.
[167] M. Musiał-Kulik, J. Kasperczyk, A. Smola, and P. Dobrzyński, "Double layer paclitaxel
delivery systems based on bioresorbable terpolymer with shape memory properties,"
International journal of pharmaceutics, vol. 465, no. 1-2, pp. 291-298, 2014.
[168] H. Gensler et al., "Implantable MEMS drug delivery device for cancer radiation reduction," in
2010 IEEE 23rd International Conference on Micro Electro Mechanical Systems (MEMS),
2010, pp. 23-26: IEEE.
[169] H. Gensler, R. Sheybani, P.-Y. Li, R. L. Mann, and E. Meng, "An implantable MEMS
micropump system for drug delivery in small animals," Biomedical microdevices, vol. 14, no.
3, pp. 483-496, 2012.
[170] R. Sheybani, H. Gensler, and E. Meng, "A MEMS electrochemical bellows actuator for fluid
metering applications," Biomedical microdevices, vol. 15, no. 1, pp. 37-48, 2013.
[171] A. Cobo, R. Sheybani, H. Tu, and E. Meng, "A wireless implantable micropump for chronic
drug infusion against cancer," Sensors and Actuators A: Physical, vol. 239, pp. 18-25, 2016.
[172] D. Stoeckel, A. Pelton, and T. Duerig, "Self-expanding nitinol stents: material and design
considerations," European radiology, vol. 14, no. 2, pp. 292-301, 2004.
[173] J. R. Laird et al., "Novel nitinol stent for lesions up to 24 cm in the superficial femoral and
proximal popliteal arteries: 24-month results from the TIGRIS randomized trial," Journal of
Endovascular Therapy, vol. 25, no. 1, pp. 68-78, 2018.
[174] A. Mehdizadeh, M. S. M. Ali, K. Takahata, S. Al-Sarawi, and D. Abbott, "A recoil resilient
lumen support, design, fabrication and mechanical evaluation," Journal of Micromechanics
and Microengineering, vol. 23, no. 6, p. 065001, 2013.
[175] Yong Xian Ang, Farah Afiqa Mohd Ghazali, and M. S. M. Ali, "Micromachined shape
memory alloy active stent with wireless monitoring and re-expansion features," IEEE 33rd
International Conference on Micro Electro Mechanical Systems (MEMS), 2020.
[176] G. Baer, T. Wilson, D. L. Matthews, and D. Maitland, "Shape‐memory behavior of thermally
stimulated polyurethane for medical applications," Journal of applied polymer science, vol.
103, no. 6, pp. 3882-3892, 2007.
[177] L. Xue, S. Dai, and Z. Li, "Biodegradable shape-memory block co-polymers for fast self-
expandable stents," Biomaterials, vol. 31, no. 32, pp. 8132-8140, 2010.
[178] A. Aiacoboae et al., "Applications of nanoscale drugs carriers in the treatment of chronic
diseases," in Nanostructures for Novel Therapy: Elsevier, 2017, pp. 37-55.
[179] Y. Luo, M. Dahmardeh, X. Chen, and K. Takahata, "A resonant-heating stent for wireless
endohyperthermia treatment of restenosis," Sensors and Actuators A: Physical, vol. 236, pp.
323-333, 2015.
[180] Y. Yi, J. Chen, M. Selvaraj, Y. Hsiang, and K. Takahata, "Wireless Hyperthermia Stent
System for Restenosis Treatment and Testing with Swine Model," IEEE Transactions on
Biomedical Engineering, 2019.
The published version is available at https://doi.org/10.1088/1361-6439/ab8832
34
[181] Y. Yi, J. Chen, Y. Hsiang, and K. Takahata, "Wirelessly Heating Stents via Radiofrequency
Resonance toward Enabling Endovascular Hyperthermia," Advanced Healthcare Materials, p.
1900708, 2019.
[182] A. L. Roy and K. Takahata, "A thermoresponsive electromechanical microchip for
temperature control in biomedical smart implants," in 2017 IEEE 30th International
Conference on Micro Electro Mechanical Systems (MEMS), 2017, pp. 460-463: IEEE.
[183] J. Chen, Y. Yi, M. Selvaraj, and K. Takahata, "Experimental analysis on wireless heating of
resonant stent for hyperthermia treatment of in-stent restenosis," Sensors and Actuators A:
Physical, vol. 297, p. 111527, 2019.
[184] Y. Luo, X. Chen, M. Dahmardeh, and K. Takahata, "RF-Powered stent with integrated circuit
breaker for safeguarded wireless hyperthermia treatment," Journal of Microelectromechanical
Systems, vol. 24, no. 5, pp. 1293-1302, 2015.
[185] A. D. Lantada et al., "Active annuloplasty system for mitral valve insufficiency," in
International Joint Conference on Biomedical Engineering Systems and Technologies, 2008,
pp. 59-72: Springer.
[186] E. Albers, D. Janssen, D. Ammons, and T. Doyle, "Percutaneous closure of secundum atrial
septal defects," Progress in Pediatric Cardiology, vol. 33, no. 2, pp. 115-123, 2012.
[187] S. Gordon et al., "Transcatheter Atrial Septal Defect Closure with the AMPLATZER® Atrial
Septal Occluder in 13 Dogs: Short‐and Mid‐Term Outcome," Journal of veterinary internal
medicine, vol. 23, no. 5, pp. 995-1002, 2009.
[188] O. Ferhanoglu, M. Yildirim, K. Subramanian, and A. Ben-Yakar, "A 5-mm piezo-scanning
fiber device for high speed ultrafast laser microsurgery," Biomedical Optics Express, vol. 5,
no. 7, pp. 2023-2023, 2014.
[189] Z. Shen, Y. Zhou, J. Miao, and K. Vu, "Enhanced Visualization of Fine Needles Under
Sonographic Guidance Using a MEMS Actuator," Sensors, vol. 15, no. 2, pp. 3107-3115,
2015.
[190] T. Nagoya and M. K. Kurosawa, "A micro ultrasonic scalpel with sensing function," in IEEE
Symposium on Ultrasonics, 2003, 2003, vol. 1, pp. 1070-1073: IEEE.
[191] M. Kurosawa and Y. Umehara, "A micro ultrasonic scalpel with modified stepped horn,"
Electronics and Communications in Japan, vol. 95, no. 8, pp. 44-51, 2012.
[192] A. Ruzzu, K. Bade, J. Fahrenberg, and D. Maas, "Positioning system for catheter tips based
on an active microvalve system," Journal of Micromechanics and Microengineering, vol. 8,
no. 2, p. 161, 1998.
[193] N. Fujiwara, S. Sawano, and S. Konishi, "Linear Expansion and Contraction of Paired
Pneumatic Baloon Bending Actuators toward Telescopic Motion," in 2009 IEEE 22nd
International Conference on Micro Electro Mechanical Systems, 2009, pp. 435-438: IEEE.
[194] S. Konishi, T. Fujita, K. Hattori, Y. Kono, and Y. Matsushita, "An openable artificial
intestinal tract system for the in vitro evaluation of medicines," Microsystems &
Nanoengineering, vol. 1, p. 15015, 2015.
[195] Y. Muramatsu, T. Kobayashi, and S. Konishi, "Flexible end-effector integrated with scanning
actuator and optical waveguide for endoscopic fluorescence imaging diagnosis," in 2015 28th
IEEE International Conference on Micro Electro Mechanical Systems (MEMS), 2015, pp.
166-167: IEEE.
[196] A. Alogla, F. Amalou, C. Balmer, P. Scanlan, W. Shu, and R. Reuben, "Micro-tweezers:
Design, fabrication, simulation and testing of a pneumatically actuated micro-gripper for
micromanipulation and microtactile sensing," Sensors and Actuators A: Physical, vol. 236,
pp. 394-404, 2015.
[197] H. Abidi et al., "Highly dexterous 2‐module soft robot for intra‐organ navigation in minimally
invasive surgery," The International Journal of Medical Robotics and Computer Assisted
Surgery, vol. 14, no. 1, p. e1875, 2018.
[198] A. Diodato et al., "Soft robotic manipulator for improving dexterity in minimally invasive
surgery," Surgical innovation, vol. 25, no. 1, pp. 69-76, 2018.
[199] S. Becker, T. Ranzani, S. Russo, and R. J. Wood, "Pop-up tissue retraction mechanism for
endoscopic surgery," in 2017 IEEE/RSJ International Conference on Intelligent Robots and
Systems (IROS), 2017, pp. 920-927: IEEE.
The published version is available at https://doi.org/10.1088/1361-6439/ab8832
35
[200] S. Russo, T. Ranzani, C. J. Walsh, and R. J. Wood, "An Additive Millimeter‐Scale
Fabrication Method for Soft Biocompatible Actuators and Sensors," Advanced Materials
Technologies, vol. 2, no. 10, p. 1700135, 2017.
[201] S. Wakimoto, I. Kumagai, and K. Suzumori, "Development of large intestine endoscope
changing its stiffness," in 2009 IEEE International Conference on Robotics and Biomimetics
(ROBIO), 2009, pp. 2320-2325: IEEE.
[202] I. Kumagai, S. Wakimoto, and K. Suzumori, "Development of large intestine endoscope
changing its stiffness-2nd report: Improvement of stiffness change device and insertion
experiment," in 2010 IEEE International Conference on Robotics and Biomimetics, 2010, pp.
241-246: IEEE.
[203] R. F. Surakusumah, D. E. O. Dewi, and E. Supriyanto, "Development of a half sphere
bending soft actuator for flexible bronchoscope movement," in 2014 IEEE International
Symposium on Robotics and Manufacturing Automation (ROMA), 2014, pp. 120-125: IEEE.
[204] B. Gorissen, M. De Volder, and D. Reynaerts, "Chip-on-tip endoscope incorporating a soft
robotic pneumatic bending microactuator," Biomedical microdevices, vol. 20, no. 3, p. 73,
2018.
[205] Y. Xu et al., "Design and development of a 3D scanning MEMS OCT probe using a novel
SiOB package assembly," Journal of Micromechanics and Microengineering, vol. 18, no. 12,
p. 125005, 2008.
[206] J. Sun et al., "3D in vivo optical coherence tomography based on a low-voltage, large-scan-
range 2D MEMS mirror," Optics Express, vol. 18, no. 12, pp. 12065-12075, 2010.
[207] L. Liu, E. Wang, X. Zhang, W. Liang, X. Li, and H. Xie, "MEMS-based 3D confocal
scanning microendoscope using MEMS scanners for both lateral and axial scan," Sensors and
Actuators A: Physical, vol. 215, pp. 89-95, 2014.
[208] S. M. H. Jayhooni et al., "Side‐Viewing Endoscopic Raman Spectroscopy for
Angle‐Resolved Analysis of Luminal Organs," Advanced Materials Technologies, 2019.
[209] S. M. H. J. I. m. S. B. Asadsangabi, H. Zeng, K. Takahata, "Ferrofluid-enabled micro rotary-
linear actuator for endoscopic three-dimensional imaging and spectroscopy," Smart Materials
and Structures, vol. 29, no. 1, p. 015025, 2020.
[210] B. A. S. M. H. Jayhooni, G. Hohert, P. Lane, H. Zeng, K. Takahata, "High-Speed and
Stepping MEMS Rotary Actuator for Multimodal, 360° Side-Viewing Endoscopic Probes," in
IEEE 33rd International Conference on Micro Electro Mechanical Systems (MEMS), 2020.
[211] H. Banerjee, M. Suhail, and H. Ren, "Hydrogel actuators and sensors for biomedical soft
robots: Brief overview with impending challenges," Biomimetics, vol. 3, no. 3, p. 15, 2018.
[212] J. Huber, N. Fleck, and M. Ashby, "The selection of mechanical actuators based on
performance indices," Proceedings of the Royal Society of London. Series A: Mathematical,
physical and engineering sciences, vol. 453, no. 1965, pp. 2185-2205, 1997.
[213] R. Pelrine, R. Kornbluh, J. Joseph, R. Heydt, Q. Pei, and S. Chiba, "High-field deformation of
elastomeric dielectrics for actuators," Materials Science and Engineering: C, vol. 11, no. 2,
pp. 89-100, 2000.
[214] H. Tobushi, S. Hayashi, and S. Kojima, "Mechanical properties of shape memory polymer of
polyurethane series: basic characteristics of stress-strain-temperature relationship," JSME
international journal. Ser. 1, Solid mechanics, strength of materials, vol. 35, no. 3, pp. 296-
302, 1992.
[215] E. Acome et al., "Hydraulically amplified self-healing electrostatic actuators with muscle-like
performance," Science, vol. 359, no. 6371, pp. 61-65, 2018.
[216] S. Ogden, L. Klintberg, G. Thornell, K. Hjort, and R. Bodén, "Review on miniaturized
paraffin phase change actuators, valves, and pumps," Microfluidics and nanofluidics, vol. 17,
no. 1, pp. 53-71, 2014.
... An implantable sensor capable of monitoring the size of the cross-section of the aneurysm could serve as an alternative monitoring approach for post-EVAR patients. Various types of wireless implantable medical sensors (WIMSs) have been proposed to cater to diverse clinical requirements for optical, neurological, and cardiovascular conditions [19,20]. Moreover, studies indicate that stent grafts can be adapted to function as intelligent sensors tailored to specific applications, such as detecting blood clots or blood flow [21][22][23][24]. ...
... Many of these devices utilise wireless inductive coupling for sensor power supply and data recording of essential physiological parameters. The development of such implantable sensors offers real-time monitoring capabilities, enabling medical professionals to remotely monitor patients from any location and without restricting their activities [19,20,25]. This remote monitoring feature allows healthcare teams to oversee patients without the necessity of them visiting medical facilities. ...
... Wireless inductive coupling systems function based on the concept that energy can be transferred between an external inductor and an implantable inductor (sensor) when they are brought into proximity [26]. Research conducted over the past few decades has focused on optimising and maximising these energy transfer systems and their associated features [19,20,25,27]. These investigated features pertain to the electrical properties of the inductors within the system, including self-inductance, resistance, parasitic capacitance, and quality factor [27]. ...
Article
Full-text available
Abdominal aortic aneurysm (AAA) is a dilation of the aorta artery larger than its normal diameter (>3 cm). Endovascular aneurysm repair (EVAR) is a minimally invasive treatment option that involves the placement of a graft in the aneurysmal portion of the aorta artery. This treatment requires multiple follow-ups with medical imaging, which is expensive, time-consuming, and resource-demanding for healthcare systems. An alternative solution is the use of wireless implantable sensors (WIMSs) to monitor the growth of the aneurysm. A WIMS capable of monitoring aneurysm size longitudinally could serve as an alternative monitoring approach for post-EVAR patients. This study has developed and characterised a three-coil inductive read-out system to detect variations in the resonance frequency of the novel Z-shaped WIMS implanted within the AAA sac. Specifically, the spacing between the transmitter and the repeater inductors was optimised to maximise the detection of the sensor by the transmitter inductor. Moreover, an experimental evaluation was also performed for different orientations of the transmitter coil with reference to the WIMS. Finally, the FDA-approved material nitinol was used to develop the WIMS, the transmitter, and repeater inductors as a proof of concept for further studies. The findings of the characterisation from the air medium suggest that the read-out system can detect the WIMS up to 5 cm, regardless of the orientation of the Z-shape WIMS, with the detection range increasing as the orientation approaches 0°. This study provides sufficient evidence that the proposed WIMS and the read-out system can be used for AAA expansion over time.
... The micro-electro-mechanical systems (MEMS) are applicable in many situations due to their advantages, such as small dimensions, low energy consumption, and long lifetime. Specific examples of such systems include semiconducting capacitors (Ye et al., 2023), radiofrequency microelectromechanical system (RF-MEM) metal-contact switch (Emhemmed and Aburwein, 2013), solar energy harvesters (Ram et al., 2022), temperature-responsive actuators (Ghazali et al., 2020;Rahaeifard, 2016;Sachyani et al., 2017), microsensors (Kong, 2013), among others. ...
Article
Full-text available
Microscale beam-like structures are standard components of micromechan-ical systems in many devices. However, the small dimensionality affects their deformation characteristics and leads to misinterpretations of the results. Presenting such size dependency is possible with advanced continuum descriptions via introducing additional variables compared to the classical one. Hence, the problems where contact occurs require the readjustment of boundary conditions considering these extra parameters. That burdens the already challenging task of contact problem calculations and restricts most demonstration examples to two-dimensional problems and geometrical linearity. The resolution of the imposed restrictions within finite element modeling further emphasizes the usage of advanced media in design and facilitates its application to various problems, which is the aim of this paper. The delivery of that task is as follows. Firstly, we present the kinematics and material descriptions for the micropolar media. The authors propose to use a newly developed continuum-based micropolar beam formulation to avoid an overwhelming computational burden and, at the same time, deal with nonlinear stress-strain relations. Secondly, the work develops a contact approach within the micropolar theory from two-dimensional to three-dimensional elasticity, although the contact is considered frictionless. Finally, it compares two existing contact formulations, including the developed one, for the contact beam problems, using the examples of two collinear sliding beams' bending.
... Micrometer-sized actuators, especially those based on microelectromechanical systems (MEMSs), have found extensive applications in various fields, such as biomedical devices, chemical sensors, CMOS-MEMS sensors, optical sensors, and switches, due to their miniature size and capabilities [1][2][3][4][5][6][7]. One of the preferred types of MEMS actuators [8][9][10][11][12][13][14] is the electrostatic actuator, known for its simplicity in terms of fabrication and its ability to generate significant force without requiring a high steady-state current. ...
Article
Full-text available
MEMS electrostatic actuators can suffer from a high control voltage and a limited displacement range, which are made more prevalent by the pull-in effect. This study explores a tri-electrode topology to enable a reduction in the control voltage and explores the effect of various solid materials forming the space between the two underlying stationary electrodes. Employing solid dielectric material simplifies fabrication and can reduce the bottom primary electrode’s fixed voltage. Through numerical analysis, different materials were examined to assess their impact. The results indicate that the primary electrode’s fixed voltage can be reduced with an increase in the dielectric constant, however, with the consequence of reduced benefit to control voltage reduction. Additionally, charge analysis was conducted to compare the actuator’s performance using air as the gap-spacing material versus solid materials, from the perspective of energy conservation. It was found that solid materials result in a higher accumulated charge, reducing the need for a high fixed voltage.
... Recent advancements in smart and portable systems demand miniaturized systems offering numerous functionality such as high speed, high resolution, ultra efficiency in performance, and most importantly, all these are achieved at the lowest possible cost [1][2][3][4][5]. MEMSs have brought revolutionary changes in the design of efficient systems, whether it is in the field of communication, sensing [6,7], biomedicals [8][9][10], safety instruments [11], food units [12], car manufacturing units [13,14], or sensor technologies. MEMS switches SMA-based MEMS switches, leveraging the unique properties of Shape Memory Alloys (SMAs), represent an innovative approach to switch technology [16]. ...
Article
Full-text available
Micro-Electro-Mechanical System (MEMS) switches have emerged as pivotal components in the realm of miniature electronic devices, promising unprecedented advancements in size, power consumption, and versatility. This literature review paper meticulously examines the key issues and challenges encountered in the development and application of MEMS switches. The comprehensive survey encompasses critical aspects such as material selection, fabrication intricacies, performance metrics including switching time and reliability, and the impact of these switches on diverse technological domains. The review critically analyzes the influence of design parameters, actuation mechanisms, and material properties on the performance of MEMS switches. Additionally, it explores recent advancements, breakthroughs, and innovative solutions proposed by researchers to address these challenges. The synthesis of the existing literature not only elucidates the current state of MEMS switch technology but also paves the way for future research avenues. The findings presented herein serve as a valuable resource for researchers, engineers, and technologists engaged in advancing MEMS switch technology, offering insights into the current landscape and guiding future endeavors in this rapidly evolving field.
... Microactuators, with their capability to precisely control and manipulate small-scale objects and systems, enable the development of miniature microelectromechanical systems (MEMS) 1-3 for applications in miniature robotics 4,5 , biomedical devices 6 and integrated electronics 7 . Magnetically responsive microactuators have been widely embraced due to their facile fabrication process, rapid response times and large range of motion 1,[8][9][10] . ...
Article
Full-text available
Microactuators provide controllable driving forces for precise positioning, manipulation and operation at the microscale. Development of microactuators using active materials is often hampered by their fabrication complexity and limited motion at small scales. Here we report light-fuelled artificial goosebumps to actuate passive microstructures, inspired by the natural reaction of hair bristling (piloerection) on biological skin. We use light-responsive liquid crystal elastomers as the responsive artificial skin to move three-dimensionally printed passive polymer microstructures. When exposed to a programmable femtosecond laser, the liquid crystal elastomer skin generates localized artificial goosebumps, resulting in precise actuation of the surrounding microstructures. Such microactuation can tilt micro-mirrors for the controlled manipulation of light reflection and disassemble capillary-force-induced self-assembled microstructures globally and locally. We demonstrate the potential application of the proposed microactuation system for information storage. This methodology provides precise, localized and controllable manipulation of microstructures, opening new possibilities for the development of programmable micromachines.
Article
Full-text available
A micro-electromechanical system (MEMS) is a micro device or system that utilizes large-scale integrated circuit manufacturing technology and microfabrication technology to integrate microsensors, micro-actuators, microstructures, signal processing and control circuits, power supplies, and communication interfaces into one or more chips [...]
Article
Full-text available
Conventional MEMS microactuators have, in recent years, been complemented by 3D‐printed actuatable microstructures fabricated via two‐Photon‐Polymerization (2PP). Herein, a novel compact 3D‐printed magnetically actuatable microactuator with a diameter of 500µm is demonstrated, originally designed for micro‐optical systems. It is fabricated by incorporating a composite of NdFeB microparticles and epoxy resin into a designated reservoir of the printed mechanical structure within a simple post‐processing step. The microactuator structure features mechanical springs, allowing for continuous positioning with large displacement. Mechanical studies by nanoindentation of IP‐S bulk structures reveal a viscoelastic material behavior, described by a two‐element General Kelvin‐Voigt viscoelasticity model. The obtained material parameters are then used to simulate and characterize the spring behavior of the microactuator. Actuation experiments are conducted using an external microcoil. The actuator displacement is measured for triangular current pulses with a peak current of 106 mA and durations of 1 to 100 s, resulting in displacements of 69.1 to 88.9 µm. Hysteretic behavior of the actuator is observed, attributable to viscoelasticity and magnetic properties of the core material. Numerical simulations of the experiment demonstrate this behavior as well. On‐the‐fly demagnetization and the implementation of closed‐loop control allow for both high repeatability and precise positioning.
Article
Full-text available
Piezoelectric accurate actuation plays an important role in industrial applications. The intrinsic frequency of previous actuators is invariable. However, variable frequency can approach the range near the low-intrinsic-frequency and realize a high actuation capability. The frequency-variable linear and rotary motion (FVLRM) principle is proposed for rotor-blade-based two-degree-of-freedom driving. Inertial force is generated by frequency-variable piezoelectric oscillators (FVPO), the base frequency and vibration modes of which are adjustable by the changeable mass and position of the mass block. The variable-frequency principle of FVPO and the FVLRM are recognized and verified by the simulations and experiments, respectively. The experiments show that the FVLRM prototype moves the fastest when the mass block is placed at the farthest position and the prototype is at the second-order intrinsic frequencies of 42 Hz and 43 Hz, achieving a linear motion of 3.52 mm/s and a rotary motion of 286.9 mrad/s. The actuator adopts a lower operating frequency of less than 60 Hz and has the function of adjusting the natural frequency. It can achieve linear and rotational motion with a larger working stroke with 140 mm linear movement and 360° rotation.
Article
Full-text available
There are many medical applications requiring angle‐resolved analysis of luminal organs. One example is lung cancer detection. Early diagnosis is the most effective way to tackle this disease. This, however, remains a challenge due to the lack of accurate detection technology. The problem is exacerbated in cases of peripheral lung cancer growing on narrow bronchi that are difficult to probe. Here, an endoscopic Raman spectroscopy device is demonstrated with a side‐viewing functionality that enables circumferential scanning spectral measurements inside thin conduits. A tubular micro rotary stepping actuator is custom‐designed and integrated with a Raman probe, for the first time, to scan a probing laser beam sideways for angle‐resolved local Raman analysis with no aid of tissue labeling, toward enabling detection of lesion‐induced biochemical changes in vivo and in real time. A microfabricated prototype is evaluated using test chemicals, harvested animal lung tissue ex vivo, as well as a murine colon model in situ and human skin in vivo. All the test results show excellent agreements with reported reference data while revealing >99% wavenumber accuracies. The study indicates that microactuator‐assisted endoscopic Raman spectroscopy is a promising technology for luminal tissue analysis and encourages further studies of its sensitivity in lesion detection.
Article
Full-text available
Thermal therapy known as hyperthermia has served as an effective method for cancer treatment. This therapeutic approach has also been attracting attention for treatment of in‐stent restenosis, the most common complication of stenting. Mild heating of stents has been shown to be a possible path to addressing this problem. Despite various studies on stent‐based thermotherapy, this area still lacks a clinically viable method and technology. Here, a radiofrequency‐powered “hot” stent prototype is reported in vitro and in vivo. An implantable stent device based on medical‐grade stainless steel acts as an electrical resonator, or an efficient wireless heater operating only when resonated using tuned external electromagnetic fields. The system architecture uses a custom‐developed power transmitter for wireless resonant powering/heating of the stent. An eight‐shaped antenna is shown to be highly effective for near‐field power transfer to the device and potentially to other smart implants, revealing stent heating efficiencies of up to 120 °C W−1, 206% of the level provided by a conventional loop antenna. Testing with swine models, the prototyped system achieves stent heating in blood flow by powering through air and skin tissue in vivo in a fully controlled manner. The results advance stent hyperthermia technology toward possible future clinical application. Endovascular hyperthermia therapy shows promising potential in suppressing in‐stent restenosis, one of the most common poststenting complications. Here, a radiofrequency‐powered resonant “hot” stent system is developed to enable stent‐based thermal treatment. Wireless heating of the resonant stent, excited through the power transmitter based on an eight‐shaped antenna, is demonstrated using swine models in vivo in a fully controlled manner.
Article
Full-text available
This paper presents a novel monolitic polydimethylsiloxane-based dual-channel bellows-structured pneumatic actuator, fabricated through a sacrificial molding technique. A finite element analysis was performed to find the optimal structure and analyze the bending performance of the square-bellows actuator. The actuator was designed with an overall cross-sectional area of 5 × 5 mm2, while its structural parameters were characterized in terms of 8, 12, 16 and 20 number of bellows. The actuator models were fabricated using acrylonitrile butadiene styrene-based sacrificial molds to form the channel and bellows structures. The experimental validation has revealed that the actuator having highest number of bellows with a size of 5 × 5 × 68.4 mm3 attained a smooth bi-directional bending motion, with maximum angles of −65° and 75°, and force of 0.166 and 0.221 N under left and right channel actuation, respectively, at 100 kPa pressure. Hence, the state-of-the-art dual-channel square-bellows actuator was able to achieve an optimal bi-directional bending with slight variations to change in temperature, which would push the boundaries of soft robotics toward the development of safer and more flexible robotic surgical tools.
Article
Full-text available
An electromagnetic (EM) micro-actuator with silicon membrane has been fabricated and characterized. The studied silicon based membrane is used as an actuator of a micropump system driven by magnetic force. The actuator consists of two main parts, namely, the electromagnetic part that generates electromagnetic field and the magneto mechanical part that enables the membrane deformation depending on the magnetic force strength on the silicon membrane. A standard Micro Electronic Mechanical System (MEMS) process was implemented to fabricate the actuator with an additional bonding between the actuator membrane and electromagnetic coil. The measurement results show that the 20 μm thin silicone membrane is capable of deformation with a maximum membrane deflection of approximately 4.5 μm which will be useful for a reliable fluids pump in a continuous drug delivery system.
Article
Various non-resonant piezoelectric actuators have been emerged in recent decades and they have been applied successfully in many industrial fields, such as medical apparatuses, optical instruments and semiconductor technologies. They are generally classified into three categories according to their different working principles: the direct drive type, the inertial drive type and the inchworm type, respectively. The main contribution of this work is to provide a comprehensive review of recent studies on the non-resonant type piezoelectric actuators. In this review, the operating principles, some representative designs, performance analyses and practical applications of each type of non-resonant piezoelectric actuators are provided. Finally, a summary is concluded and the future development perspectives of these non-resonant type piezoelectric actuators are discussed. This review contributes to giving a comprehensive understanding of the non-resonant type piezoelectric actuators. It also provides guidance on designing new piezoelectric actuators for improving their mechanical output performances.
Article
In this paper, an impulsion system of liquids for thermoplastic microfluidic circuits is described. The presented device is composed of a membrane made of polymethylmethacrylate (PMMA) which separates two microchambers (top and bottom microchamber). The bottom microchamber is pressurized to a fixed pressure, while the top microchamber remains at atmospheric pressure. The actuation consists on increasing the temperature of the membrane by Joule effect using an aluminum microheater which is fabricated over the membrane. Once the temperature of the membrane is close to the glass transition temperature of the thermoplastic, the membrane deforms due to the different pressure of the bottom microchamber, blocking the top microchamber. The trapped air fits the volume of liquid which will be impulsed inside the microchannel. Unlike other membranes used in microfluidics, the presented membrane keeps deformed when the actuation is removed. The fabricated impulsion system is designed to move a volume of 30 μL. The experimental results correspond to a theoretical design with an average error of about 6.32%. The proposed impulsion system is easily integrable on thermoplastic lab on chips for liquid management.