Next Article in Journal
Structural and Thermal Stability of CrZrON Coatings Synthesized via Reactive Magnetron Sputtering
Previous Article in Journal
The Application of Tribological Tests to Study the Wear Behavior of Asphalt Cutter Teeth: An Experimental Investigation Using Baroid Tribometer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review of Flexible Acceleration Sensors Based on Piezoelectric Materials: Performance Characterization, Parametric Analysis, Frontier Technologies, and Applications

1
School of Applied Science, Beijing Information Science & Technology University, Beijing 100192, China
2
Key Laboratory of Sensors, Beijing Information Science & Technology University, Beijing 100192, China
3
Key Laboratory of Modern Measurement & Control Technology, Ministry of Education, Beijing Information Science & Technology University, Beijing 100192, China
4
Department of Electrical and Computer Engineering, Carnegie Mellon University, Pittsburgh, PA 15213, USA
5
State Key Laboratory of New Ceramics and Fine Processing, School of Material Science and Engineering, Tsinghua University, Beijing 100084, China
*
Authors to whom correspondence should be addressed.
Coatings 2023, 13(7), 1252; https://doi.org/10.3390/coatings13071252
Submission received: 19 May 2023 / Revised: 6 July 2023 / Accepted: 13 July 2023 / Published: 15 July 2023
(This article belongs to the Section Thin Films)

Abstract

:
Acceleration sensors are tools for detecting acceleration and serve purposes like fault monitoring and behavior recognition. It is extensively employed in a variety of industries, including aerospace, artificial intelligence, biology, and many more. Among these, one of the major research hotspots and challenges is the development of low-energy, self-powered, miniature, mass-produced sensors. Due to its capacity to perceive human behavior and identify errors, the flexible acceleration sensor offers a distinct advantage in the use of flexible and miniaturized sensing systems. This review analyzes the current state of piezoelectric flexible acceleration sensors’ applications in the areas of sensitive materials, processing technology, and device structure and briefly summarizes the fundamental properties of these sensors. Additionally, it ends with a prognosis for the future growth of flexible piezoelectric acceleration sensors.

1. Introduction

Sensing technology is an important support in the field of modern information technology, and together with communication and computers, it can collect, transmit, and process external measurement signals [1,2]. An acceleration sensor is a device that senses acceleration, which allows the acquisition of various vibration signals to further perform functions such as behavior recognition and fault monitoring [3,4]. It has important applications in aerospace, power and energy, power equipment, health monitoring, and other fields [5]. Driven by the new generation of information and communication network technology, the industry and Internet will develop in a broader, deeper, and higher direction [6,7]. The manufacturing industry will rapidly shift to digitalization, networking, and intelligence [8], and perception is a key source of information for this process. Flexible perception is an important technological support for achieving multi-dimensional spatial freedom [9]. Flexible sensing can change the traditional physical form of information devices and systems to organically integrate information with people, objects, and environments [10]. The flexibility of acceleration sensors is therefore also an important upgrade for functions such as behavior recognition and fault monitoring. Acceleration sensors can be divided into capacitance types, resistance types, optical fiber types, inductance types, and electric vortex types [11,12]. Among them, piezoelectric material is used as the sensitive material for piezoelectric acceleration sensors. Piezoelectric materials can realize the mutual conversion between vibration signals and electrical signals to directly generate electrical signals in response to vibration signals without external power supplies, which can realize a zero-power consumption system design [13,14] and has a unique advantage in the application of flexible and miniaturized sensing systems.
In this review, the most recent advances in the manufacturing technology, device structure, and sensing material research of flexible piezoelectric acceleration sensors are summarized. It is also introduced simultaneously along with the performance parameters. The potential future of the flexible piezoelectric acceleration sensor is briefly discussed in Section 6.

2. Development, Parameters, and Relationship between Parameters of Flexible Piezoelectric Acceleration Sensors

2.1. Development

Flexible piezoelectric acceleration sensors are devices with both flexible and acceleration-sensing functions, and their core sensitive materials are piezoelectric materials. Flexible piezoelectric acceleration sensors have broad application areas in behavior recognition, robotics, and the Internet of Things (IoT), as shown in Figure 1 [15,16,17,18,19,20,21,22,23,24,25,26,27].
When compared to a standard acceleration sensor and a single flexible acceleration sensor, the flexible piezoelectric acceleration sensor is superior; it can be considered a combination and further development of both. It can achieve self-power supply, high stability, and high toughness based on the benefits of piezoelectric materials and lacks the awkwardness and low sensitivity of conventional accelerometers. Figure 2 shows the comparison of the above sensors and the common structure of the flexible piezoelectric acceleration sensor [28,29,30,31,32,33,34,35].

2.2. Parameters

The knowledge of basic parameters is critical for the selection of pressure sensors as every application scenario has its unique requirements for sensor characteristics [36]. The performance of flexible piezoelectric accelerometers is evaluated by the following parameters.

2.2.1. Frequency

The natural frequency is the frequency that causes the sensor to exhibit the resonance phenomenon. The external vibration frequency forces the sensor to generate spontaneous vibration, and when the vibration amplitude reaches its maximum, it is resonant. Resonance will generate heat in the sensor heat, as well as structural damage [37], so it is necessary to use the sensor below this frequency. The natural frequency (resonance frequency) is given by
f = k m 2 π
where m is the mass of the sensor and k is the structural stiffness.

2.2.2. Sensitivity

Sensitivity is the most basic and important parameter of an accelerometer. The sensitivity of a piezoelectric sensor is defined as the ratio of electrical output to mechanical input. Higher measurement accuracy can be obtained by increasing the sensitivity. For piezoelectric vibration sensors, charge sensitivity can better characterize the properties.
It is known that the acceleration value is proportional to the square of the signal frequency for the same displacement, so the amplitude of acceleration signals in different frequency ranges can vary considerably and may even exceed the measurement range of the accelerometer, causing distortion. As can be seen, the sensitivity of the accelerometer needs to be fully estimated when measuring high-frequency signals and low-frequency signals. As a general principle, the higher the sensitivity, the smaller the measuring range, and conversely, the smaller the sensitivity, the larger the measuring range.

2.2.3. Pre-Tightening Torque

The preload force is the force that is added in advance to enhance the reliability and tightness of the connection before the relevant structure is subjected to the working load, and also to prevent gaps and relative slippage between the connected devices after being subjected to the load. The magnitude of the preload force has an important influence on the degree of stiffness of the part, and therefore the preload force is also one of the main factors affecting the resonant frequency [38].
If the pre-tightening force is small, the stiffness of the sensor will decline, resulting in a lower resonant frequency [39], which ultimately increases the product’s upper limit of frequency response and frequency response error. As the pre-tightening force gradually increases, the product stiffness, resonant frequency, and the upper limit of frequency response will increase, but the error will decrease. Of course, the increase in pre-tightening force is limited, and when it approaches the bearing limit of the part structure and sensitive elements, it will lead to the structure’s deformation or even damage, causing reductions in stiffness and frequency response limit and an increase in error [40].

2.2.4. Noise

Noise is the interference signal superimposed on the useful signal, which has a large interference with the analysis and utilization of the measured signal. The noise of the measurement system can be divided into two parts: internal noise and external noise. The signal interference outside the measurement system itself, such as the vibration of the external environment and the interference of the electromagnetic field, is the main source of the external noise. Additional shielding measures can be used to achieve the purpose of noise reduction for such signals.
Internal noise comes from the system itself, also known as inherent noise or local noise, and can only be reduced by designing reasonable parameters. Internal noise determines the resolution and limits the dynamic measurement range of the system. For example, low-frequency vibration signals are very weak and must be measured with a low-noise vibration sensor to ensure that the useful signal is not submerged by the background noise. The internal noise of piezoelectric accelerometers can be further divided into mechanical thermal noise, electrical thermal noise, and noise generated by charge amplifiers [39]. Mechanical thermal noise is related to the mass, spring coefficient, and mechanical resistance of the sensor inertial system. It can be reduced by increasing the mass and quality factors or reducing the resonant frequency. Electrical thermal noise is related to the loss factor of sensing materials, and studies have shown that materials with fewer impurities and defects are often used for low-noise devices [41]. In the study of sensors, 1/f noise and its thermal noise are mainly considered in the noise generated by charge amplifiers [42]. In addition, the sensor can sometimes be regarded as a capacitor and the presence of capacitors in charge amplifiers can also lead to the generation of KT/C noise [43], which cannot be ignored.

2.3. Relationship between Parameters

As shown in Figure 3a,b [44,45], the output voltage and sensitivity of the piezoelectric accelerometer are plotted against frequency, respectively. There is an inverse ratio between sensitivity and resonant frequency, and only at the expense of one parameter can the other be improved. If the sensor has a higher sensitivity, smaller changes in the measurement can be recognized. But a small interference in the measured object can lead to a significant change in the output of the sensor. When a small disturbance occurs, the signal is quickly amplified, which greatly affects the stability and detection range of the sensor. Therefore, by increasing the sensitivity, the accuracy of the detection can be improved, but this needs to be done within reasonable limits. Figure 3c [46] shows the sensitivity of the sensor measured at different preload torques. It can be seen that within a certain range of values, increasing the preload force appropriately improves the sensitivity of the sensor and its frequency response is smoother in the lower frequency bands, because if the preload force is appropriate, the contact stiffness and pre-strain of the element will improve, leading to an increase in the stress applied to the element, induced charge generation, and increased sensitivity. And once this threshold is exceeded, the sensitivity and frequency response performance of the system will decrease. This phenomenon may be due to the deformation of small bumps on the wafer surface beyond the elastic limit, plastic deformation, or cracking, which increases the contact area and reduces the contact stiffness, resulting in reduced sensitivity and poor stability of the frequency response [47]. In Figure 3d [48], the temperature gradually increases from 0 to 1000 °C and the frequency excitation increases from 200 to 600 Hz; the sensitivity of the sensor still fluctuates around a horizontal line, reflecting the good temperature stability of the sensor in a certain range. In Figure 3d, the temperature gradually increases from room temperature to 650 °C and the frequency excitation increases from 120 to 520 Hz; the sensitivity of the sensor still fluctuates around a horizontal line, reflecting the good temperature stability of the sensor in a certain range. Figure 3e [49] shows the output voltage noise spectral density of piezoelectric acceleration sensors. It can be seen that the main noise sources are the thermal noise of the bias resistor, and the input current and voltage noise of the operational amplifier at high frequencies [50]. Figure 3f [49] shows the output charge as a function of acceleration for the 600 Hz case. The figure shows that the charge sensitivity is 2.3 pC/g at both room temperature and 900 °C, with a deviation at 900 °C because more noise disturbances are introduced at this time compared to room temperature.
Table 1 shows the performance comparison of different types of piezoelectric acceleration sensors. It should be pointed out that due to the different designs and use ranges of different acceleration sensors, their sensitivity units are also different. Obviously, the input of the acceleration sensor is acceleration, which is generally expressed by the gravitational acceleration g. If the output of the sensor is a voltage change, such as mV, then the unit of sensitivity is mV/g; if the output is a change in the amount of charge (e.g., pC), the sensitivity is measured in pC/g.

3. Materials for Flexible Piezoelectric Acceleration Sensors

Flexible substrates, electrodes, functional components, and packaging materials make up the majority of flexible electronic devices [57]. Flexible piezoelectric acceleration sensors are consistent with it, and its performance and use are typically impacted by its piezoelectric material, substrate material, preparation circumstances, and preparation procedure. The core of the smart age is made up of piezoelectric materials, which perform both active (producing sound waves and moving objects) and passive (sensing) activities. Piezoelectric materials are frequently employed to capture outside signals and are a crucial component of the intelligent era [58]. The foundation substance can support the adaptability of sensor components and further safeguard the upkeep and regular operation of the sensor form. To make the piezoelectric sensor exhibit a high response output and good flexibility, it is best to choose piezoelectric materials with a high piezoelectric coefficient and flexibility, as well as ensure sensitivity and mechanical endurance. Sometimes, there are substrates between the samples to be measured and the sensors, which affect how the two interact. In this case, it is necessary to establish a connection between the substrate and the respective electric system, and comprehensively consider solving the practical problem. Hofmann et al. invented a textile-reinforced structure that can itself be used for sensing and integrated into a fiber-reinforced plastic composite to enable structural health detection. Apparently, the fiber-reinforced plastic composite is used as a substrate for insulation. In order to connect to the electric system, the researchers used copper tape with conductive adhesive to achieve electrical contact and, after solidification, exposed the tape by removing the hardening resin and finally soldered to the cable [59]. That is, in the case of the substrate as an insulating layer, the wiring can be completed on the substrate and connected to the power system. In fact, it is also possible to connect using wireless communication methods such as Bluetooth and WiFi. When the substrate is used as a protective layer, the power system of the substrate should try to ensure that the signals received by the substrate and sensors are not distorted and attenuated, with specific measures that include shielding, noise reduction, impedance matching, and so on. Microstrip antennas for sensor signal transmission are widely studied, which can realize wireless transmission, mask the sensor without affecting its ability to receive a signal proportional to the measurement signal, and also realize a low response of the device in different scattering angle cases through scattering offset mechanisms and stealth characteristics, which will be extremely beneficial to the stability of sensor signal transmission [60].
Some material parameters can conversely be measured by piezoelectric sensors, such as the measurement of sample texture, which involves the problem of inverse scattering of the sample texture [61]. The problem is to determine the basic material properties or surface characteristics of an object from the measurement or observation of scattered waves or signals. When an incident wave or signal interacts with a textured surface, scattering occurs and a scattering signal is obtained. Using a piezoelectric sensor, the acquired signal is collected, processed, and analyzed, and the sample texture can be analyzed. However, in the specific implementation, the determination of sample texture using piezoelectric sensors is still a great challenge because of the limited data, the complexity of the calculation, and the uniqueness of the results. Advanced mathematical and computational methods need to be developed to cope with it. Stoynov et al. studied the scattering problem of cracked piezoelectric planes, and with known far-field data, the stress intensity factor (SIF) at the crack edge was found [62]. Valagiannopoulos et al. derived the dielectric coefficient of anisotropic materials using the Taylor series expansion and appropriate polarization excitation after measuring the projection coefficient [63].
However, piezoelectric materials such as ZnO, PZT, BaTiO3, and PVDF, which have been widely studied at present, have not achieved the function of having both flexibility and excellent piezoelectric properties. Although ZnO is a semiconductor, PZT, BaTiO3, and other induced ceramics have high ferroelectricity, but they are poor in flexibility. In addition, PZT-based piezoelectric materials will be gradually abandoned because of their harm to the environment and biology. Similarly, PVDF and its copolymers are flexible, but their piezoelectric properties are too poor. Crystal structures, SEM images, and related devices of four piezoelectric materials, ZnO, PZT, BaTiO3 and PVDF, were compared, as shown in Figure 4 [64,65,66,67,68,69,70,71,72,73,74,75].
To sum up, there is an urgent need to explore and design materials with high voltage and high flexibility to fabricate piezoelectric acceleration sensors [76]. It was found that piezoelectric polymers such as PVDF and PVDF-TrFE can be used for the preparation of current flexible piezoelectric acceleration sensors. The preparation methods are divided into two types: One is to directly prepare piezoelectric polymers into nanofibers by electrospinning and other technologies. For example, piezoelectric polymers can be processed into PVDF-TrFE nanofibers. Ghosh et al. realized the preparation of this material, finding good softness and air permeability of this material (see Figure 5a) [77]. Another example is that inorganic piezoelectric materials are compounded with flexible polymers as nano-fillers, and then after micro-processing, flexible piezoelectric sensors can be fabricated [78]. Ceramic/polymer composites were first proposed in 1978, in which electrical properties have been greatly improved and are expected to exhibit some new properties [79,80]. Inorganic materials in the material will be very sensitive to applied forces due to high voltage [81,82]. In addition, due to the high elasticity of the elastic matrix in the composite, they also have good shape preservation and flexibility. Common inorganic piezoelectric materials include LiNbO3, PbTiO3 nanowires, PbTiO3 nanoparticles, etc. Elastic matrix is divided into non-piezoelectric organic materials and piezoelectric organic materials, and the former mainly include polydimethylsiloxane (PDMS), polyvinyl alcohol (PVA), and polystyrene (PS). Figure 5b–d show the composite process of piezoelectric ceramics and piezoelectric organic materials, among which Figure 5b shows the overall composite process [83] and Figure 5c [84] shows the change in material microstructure during the composite process, and Figure 5d [85] further shows the networking situation of piezoelectric ceramic particles in piezoelectric organic materials of PVDF. The composite of piezoelectric ceramics and non-piezoelectric organic materials is shown in Figure 5e–h [86]. Alluri et al. synthesized a flexible film of BaTiO3 and PDMS. Finally, they found that the film had good flexibility and rolling ability. This is because PDMS is chemically stable, highly insulating, inexpensive, corrosion-resistant, resistant to high and low temperatures, has good dielectric properties, is highly elastic, and has a high recovery force. It is cross-linked into a film by a polydimethylsiloxane curing agent [87] and can be stretched to 300% of its original length [88].

4. Structure of Flexible Piezoelectric Acceleration Sensors

The traditional piezoelectric acceleration sensor structure is shown in Figure 6a [83]. According to the classification of the vibration mode and sensor structure, traditional acceleration sensors can be divided into compression-type (d33), shear-type (d15), and bending-type (d31) [89,90]. Compression-type piezoelectric acceleration sensors mainly consist of a fixation nut, a mass block, a piezoelectric sheet, and a base. The number of piezoelectric pieces can be increased or decreased, but it is necessary to pay attention to whether the series or parallel connection is used. Shear-type sensors are mainly composed of a fixed screw, a mass block, a piezoelectric piece, and a base [91]. Unlike compression sensors, which can use a single piezoelectric plate, shear sensors generally use pairs of piezoelectric plates. The structure of a flexible piezoelectric acceleration sensor is similar to that of a traditional piezoelectric acceleration sensor, but many rigid components are replaced by flexible materials to meet the need for flexibility.
Wang et al. [92] presented the design and testing of a uniaxial piezoelectric acceleration sensor made from cellulose paper and zinc oxide nanowires (ZnO-NWs). The device is shown in Figure 6b. As only the paper needs to be cut and then the growth of ZnO nanoparticles is carried out in a hydrothermal method, it does not require expensive and complex instrumentation. Figure 6c illustrates the experimental apparatus for this paper-based accelerometer with a sensitivity of 16.3 mV/g at a natural frequency of 84.75 Hz for this accelerometer weighing 61 mg. Zhang et al. [93] developed a 1D–3D fully piezoelectric nanocomposite based on perovskite BaTiO3 and PVDF-TrFE and applied it in a material generator (hNCG) device (see Figure 6d). The output of such a flexible hNCG can be as high as 14 v and 4 μA, which is significantly higher than current devices made using piezoelectric ceramic films. The finite element analysis showed that the superiority of the piezoelectric composite is due to the piezoelectric synergy of the two materials and the effective pressure transfer ability of the piezoelectric composite.
Respiratory health surveillance with very low detection limits is extremely challenging. Given this, Yuan et al. [94] from Xidian University designed a flexible bionic sensor with a detection limit as low as 0.0005 N. The specific structure of the sensor is shown in Figure 6e, which has the advantages of short response time and high response repeatability. In addition, reasonable design and optimization of the back-end circuit can make it monitor various breathing states, including apnea, cough, and deep breathing, which is extremely expandable and a breakthrough in the wearable field.
Zhang et al. [95] designed a self-powered high-sensitivity acceleration sensor based on the triboelectric nanogenerator, as shown in Figure 6f. The sensor is composed of an LMMD and an nn-PVDF film, where the LMMD has high surface tension, high mass density, high elasticity, and mechanical robustness. The measurement shows that the acceleration sensor detection ranges from 0 to 60 m/s2. It also exhibits excellent stability, having potential applications in equipment vibration measurement monitoring and troubleshooting.
Figure 6g shows the fabrication process of a flexible piezoelectric sensor made from a group III nitride film, designed by Chen et al. [96]. Group III nitrides exhibit environmentally friendly, chemically stable, biocompatible, resistant, and stable input and output characteristics.
In Figure 6h, a PVDF sensor patch [97] was designed to detect heartbeat and respiratory signals. The sensor patch is composed of three layers: the PDMS covering module, the PVDF polymer film, and the Mylar layer from top to bottom. The upper and lower surfaces of the PVDF sensor film are, respectively, equipped with two printed electrodes for signal conduction and grounding, which are connected to a shielded cable to prevent electrical noise. The Mylar layer provides insulation between the printed electrodes and the chest cavity. It was found that compared with the flat PVDF film structure, the curved structure can increase the detection signal by 151%.

5. New Technologies of Flexible Piezoelectric Acceleration Sensor

5.1. Three-Dimensional Printing Technology

5.1.1. Introduction to 3D Printing Technology

The future application areas of sensors will develop in several directions, such as real-time monitoring, advanced diagnostics, and interconnected feedback, which will undoubtedly place higher demands on the sensitivity, resolution, strength, and hardness of sensors. The emergence and development of 3D printing technology obviously provide great possibilities for this.
Three-dimensional printing technology, also known as additive manufacturing, enables the production of precisely shaped, smooth-surfaced parts, where the printing process can be controlled automatically, started and stopped at any time, or incorporated into complementary manufacturing processes or embedded in sub-components manufactured with conventional processes. The production process for 3D-printed sensors can thus be seamlessly achieved by embedding the sensor into a printed structure or by printing the entire sensor from the ground up [98]. Sensors with complicated geometry, complex functionality, and ease of assembly are becoming possible thanks to advancements in multi-process and hybrid 3D printing technologies [99,100].

5.1.2. Three-Dimensional Printing Technology Classification and Process Flow

The whole process of 3D printing technology is shown in Figure 6 [101]: First, a model of the object to be printed is created using software such as Solidworks. Once the 3D model is created, it is then converted into an STL file, which stores all the information about the details of the model. The 3D printer obtains a series of 2D cross-sectional layers by splitting in the Z-direction. Finally, the desired object is printed layer by layer. Taking the printed flexible sensor and “glove” as examples, Figure 7(bi) shows the CAD design of the flexible sensor. Figure 7(bii) and Figure 7(biii) show the un-flexed and flexed states of the sensor, respectively. Figure 7(ci) and Figure 7(cii), respectively, show the CAD design of “gloves” and the finished product after printing. Figure 7(ciii) and Figure 7(civ), respectively, show the printed “gloves” in two states: un-flexed and flexed [102]. The piezoelectric nanocomposites were printed using 3D printing technology and the device used was a custom PuSL fabrication system, as shown in Figure 7d. As seen in Figure 7e–h, the printed piezoelectric nanocomposites have a fine surface structure and excellent surface finish [103]. Bodkhe et al. reasonably adjusted the nozzle size, pressure, and solution concentration to obtain 1D, 2D, 2.5D, and 3D shapes of PVDF, as shown in Figure 7i–l [104].
Three-dimensional printing technologies can be subdivided into seven main categories according to specific manufacturing techniques, and their detailed technical processes and characteristics are as follows:
  • Fused Deposition Modeling (FDM)
The process works by melting and extruding a thermoplastic filament from a nozzle, which is mainly from thermoplastic materials such as polyamide (PA), polylactic acid (PLA), and polycarbonate (PC). The melted material is deposited on a production platform, where it is cooled and solidified to form a single layer. This process is repeated layer by layer to form a 3D model. FDM has the characteristics of being low cost and open source, but its print resolution and print speed are often low. Crump first proposed FDM [105], its device is shown in Figure 8a [106].
Based on the thermoplastic properties of PVDF and taking advantage of the combination of FDM and the corona poling process (IPC), Kim et al. achieved the low-cost, low-complexity, and flexible fabrication of PVDF piezoelectric sensors [107]. In the same year, they also found that FDM could significantly improve the uniform dispersion of BT nanoparticles in PVDF matrix, enhancing the piezoelectric properties. As a result, they designed a Multiwall carbon nanotube (MWCNT)/BT/PVDF nanocomposite film, whose tests showed excellent mechanical properties and durability [108].
2.
Direct Ink Writing (DIW)
Direct ink writing, as shown in Figure 8b [109], uses nozzles to extrude materials directly onto the manufacturing platform, where the material is deposited in a viscous liquid state to maintain its shape after deposition. Direct inkjet technology can be used in many materials, such as ceramics, plastics, food, hydrogels, living cells, and so on. The size of the nozzle, the viscosity and density of the material, the scanning speed, the jet speed, and other parameters can all be adjusted to obtain an optimal deposition object. Post-processing may require hardening the manufactured object and improving its mechanical properties through sintering, heating, UV curing, and drying.
DIW technology is suitable for multi-material, free-form, complex geometry printing and is often used for the one-step printing of force sensors. Using DIW to print composites of PDMS and barium titanate (BTO), the resulting samples are high-density, have a high-voltage electrical coefficient, and have tunable mechanical properties; however, the prints have a low resolution and are subject to complex post-processing processes such as high-temperature curing and photopolymerization [110].
3.
Photocuring (SLA, DLP)
Photocuring is curing layer by layer by UV light to form a three-dimensional structure on the platform. Light curing technology can be divided into two types according to the light source and the way it is processed: stereolithography equipment (SLA) and digital light processing (DLP) [111]. The former uses a moving laser to cure the photosensitive resin directly, as shown in Figure 8c [112], while the latter uses a laser or UV lamp as a light source to cure the exposed parts by means of a special pattern. SLA and DLP can produce high-precision structures with complex internal characteristics but have one disadvantage: only a single material can be used.
Biosensors require extremely complex information to be processed in a small mechanical building block, but conventional fabrication methods often struggle to achieve extremely strong mechanical coupling within the sensor, which makes it difficult to apply biosensors for high-precision measurements on tiny objects. Tiller et al. [113] proposed the use of DLP to achieve this requirement. First, barium titanate nanopowders were mixed with multi-walled carbon nanotubes (MWCNTs) into a PEGDA-based monomer resin, and the MWCNTs were oriented and aligned by an electric field to make piezoelectric devices.
4.
Lamination (LOM)
Figure 8d [114] shows the experimental apparatus designed for the LOM. The additive manufacturing technology for LOM was first developed by Helisys in 1988 [115]. It first uses a laser or knife to cut the sheet material, and after cutting off a layer, a new layer is firmly adhered up by a roller that compacts and heats/glues the sheets together. This process is repeated and the unwanted parts are later removed to obtain a complete 3D structure [116].
5.
Selective Laser Sintering and Selective Laser Melting (SLS and SLM, respectively)
Both SLS and SLM are based on powders, including plastics, metals, ceramics, and waxes, which can be printed with sufficient strength and density to meet aerospace or military requirements [117]. The device for SLA is illustrated in Figure 8e [118].
Song et al. [119] showed that SLS can be used to print PVDF/graphene nanocomposites with hollow axial-shaped arrays resulting in parts with higher conversion efficiency, higher piezoelectric performance with an optimum open-circuit voltage of 16.97 V, a short-circuit current of 274 nA, and the ability to light up to 10 LED bulbs simultaneously.
6.
Photopolymer Jetting (Ployjet)
The photopolymer injection technique was first proposed by Gothait [120]. The printing material chosen was a photosensitive resin, which was squeezed out of a nozzle and deposited onto a printing platform, and then it was cured by UV light, as shown in Figure 8f [110]. This method also requires layer-by-layer manufacturing. The advantage of optical polymer injection is that it can print a variety of materials and colors at the same time, and it has a high resolution, which is suitable for printing small and delicate objects. The disadvantage is that the parts produced by this method are weaker.
Ployjet enables printed, highly integrated sensors and actuators that are constructed from a variety of photosensitive polymers [121]. The ployjet-based 3D printing technology can also be used to prepare multifunctional, highly integrated, and efficient microfluidic devices. Based on previous research, Nigtingale et al. developed a microfluidic-chip-based wearable bioassay system for monitoring biomolecular levels using liquid chromatography technology [122].
Figure 8. Schematic diagram of the equipment of 3D printing. (a) Fused deposition modeling (FDM). Reprinted with permission from Ref. [106]. Copyright 2018, MDPI AG. (b) Direct ink writing (DIW). Reprinted with permission from Ref. [109]. Copyright 2020, Elsevier. (c) Stereolithography (SLA). Reprinted with permission from Ref. [112]. Copyright 2014, American Chemical Society. (d) Lamination (LOM). Reprinted with permission from Ref. [114]. Copyright 2016, John Wiley and Sons. (e) Selective laser sintering (SLS). Reprinted with permission from Ref. [118]. Copyright 2018, American Chemical Society. (f) Photopolymer jetting (Ployjet). Reprinted with permission from Ref. [109]. Copyright 2017, MDPI AG. (g) Binder jetting (3DP). Reprinted with permission from Ref. [123]. Copyright 2018, MDPI AG.
Figure 8. Schematic diagram of the equipment of 3D printing. (a) Fused deposition modeling (FDM). Reprinted with permission from Ref. [106]. Copyright 2018, MDPI AG. (b) Direct ink writing (DIW). Reprinted with permission from Ref. [109]. Copyright 2020, Elsevier. (c) Stereolithography (SLA). Reprinted with permission from Ref. [112]. Copyright 2014, American Chemical Society. (d) Lamination (LOM). Reprinted with permission from Ref. [114]. Copyright 2016, John Wiley and Sons. (e) Selective laser sintering (SLS). Reprinted with permission from Ref. [118]. Copyright 2018, American Chemical Society. (f) Photopolymer jetting (Ployjet). Reprinted with permission from Ref. [109]. Copyright 2017, MDPI AG. (g) Binder jetting (3DP). Reprinted with permission from Ref. [123]. Copyright 2018, MDPI AG.
Coatings 13 01252 g008
7.
Binder Jetting (3DP)
This method is similar to the aforementioned method, except that this method does not require any support structure, as the powder can support itself. Powder materials are often ceramics, gypsum, sugar, etc. Adhesive jetting is mainly used in the biological field, and it can print a wide range of materials, but the strength and surface roughness of the object are not so good. Figure 8g [123] shows the device used for 3DP.
The classification, materials used, advantages, and disadvantages of 3D printing technologies are summarized in Table 2:

5.1.3. Influence of 3D Printing Parameters on Sensors

The benefits of 3D-printed piezoelectric acceleration sensors are evident to all, but for this sensor to have high performance, it often requires good tuning of the parameters associated with 3D printing, which include extrusion rate, nozzle diameter, nozzle temperature, and bed temperature. The extrusion rate is positively correlated with the pressure and shear rate at the narrowing of the nozzle, and its variation may cause the print trace spacing to change and overlap [124]. Smaller nozzle diameters can damage the structure of materials such as fibers or cause nozzle clogging, resulting in reduced sensing capabilities [125], while using larger nozzles may reduce the print resolution [126]. Nozzle temperature and bed temperature also need to be strictly controlled, because above the appropriate temperature range, the material viscosity will be reduced, the print traces will be deformed, and even the material will be degraded [127]; below the appropriate temperature range, the material will not adhere to the surface of the printing bed. Kim et al., while using the 3D printing process to fabricate PVDF films, found that too high a bed temperature reduces the piezoelectric properties of PVDF films and decreases the sensitivity [108].

5.1.4. Application of 3D Printing Sensor

It is known that high-resolution physical sensors or microfluidic devices can be used for particle detection [128] and 4 μm single-cell capture [129]. However, the traditional manufacturing method is not only complicated, time-consuming, and lacks scalability and extensibility, but also has high manufacturing costs to achieve a high-resolution structure. With the rapid development of 3D printing technology, the characteristics of short manufacturing time and low manufacturing cost can obviously make up for these defects. The cost of a sensor manufactured by conventional methods is USD 215, while the cost of such a sensor printed by 3D printing (stereolithography) is limited to USD 200 [130]. In addition, microfluidic devices with complex structures, such as multiple PDMS layers, can be reproduced simply and quickly by 3D printing technology, without the need for soft lithography manufacturing requiring complicated processes such as optical rotation resistance, coating, and UV exposure, saving a lot of cost and labor.
An even greater feature of 3D-printed sensors is that they can achieve high sensitivity, such as omnidirectional radiation pattern sensing, which can be obtained by 3D printing with a larger response surface area, to achieve 2.25 times sensitivity [131,132,133,134,135]. Three-dimensionally printing sensors can also achieve high flexibility. It makes it easy to change the geometry of the sensor so that it can be paired with electrodes as an electrical sensor, or with smartphones as an optical sensor to match measurements with other commercial products of known and measurable size.
A flexible sensor typically consists of a sensing element, a flexible substrate, flexible electrodes, packaging, and wiring. The application of 3D printing to sensors varies depending on the components printed and often includes (1) custom molds with various surface structures to obtain sensor substrates, electrodes, and sensing elements; (2) the printing of flexible bases or major parts of sensors that are then used to compose sensors with good flexibility and ductility; (3) the manufacture of sensing elements with specially tailored geometries and microstructures; (4) the suitability for printing flexible electrodes. In some sensors, flexible electrodes are required to take advantage of their conductive and stretchable properties, which can lead to better impedance stability [136]; (5) complete printing of the entire sensor [120].
Figure 9a shows the design of a flexible tactile sensor proposed by Guo et al. [137]. The design develops a non-sintered ink with electrical conductivity and adjustable viscosity. Using a multi-material, multi-scale, and multi-functional 3D printing method, the designed sensor shows good linear current–voltage characteristics (see Figure 9b for the detailed process flow). Figure 9c [138] shows a piezoelectric pressure sensor fabricated by Fuh et al. The sensor substrate was fabricated by an FDM process and has a wavy shape. A layer of copper foil was placed on the substrate, and the piezoelectric composite PVDF was deposited on top of it by electrospinning. Finally, the sensor was encapsulated using PDMS film. Figure 9d [139] shows a functional acceleration sensor for piezoelectric signal readout. Bernasconi et al. combined 3D printing and inkjet material deposition intelligently to create a PVDF-TrFE piezoelectric layer and corresponding silver electrode. At the same time, stereolithographic techniques were selected to fabricate the structural components of the accelerometer. The results show that the proposed hybrid additive manufacturing technique is very promising for the mesoscale manufacture of electromechanical sensors.

5.2. MEMS Technology

5.2.1. Introduction to MEMS Technology

Prior to measurement, some sensors that need contact measurements are glued to the host structure. Obviously, this is less ideal for smaller structures due to the size of the sensor itself as well as the adhesive’s excessive thickness, which could alter the substrate’s behavior. Sensors should scale with the size of the host structure and should be optimally placed. The optimal sensor placement is a function related to the geometry of the structure and the vibration signal to be retrieved, which also requires the sensor to be much smaller than the host structure [140]. MEMS technology clearly meets these requirements.
MEMS technology gradually developed in the 1990s as an emerging technology in the field of micromechanics, which uses integrated circuit technology processes and micromachining methods to integrate electromechanical sensitive components and processing circuits based on various physical effects on a chip, successfully realizing the combination of micromachining and microelectronics technology [141]. Rather than improving the performance of these systems while reducing size and power, as traditional engineering efforts have done, the MEMS field has worked to fundamentally change the size, performance, and cost [142,143]. With the development of MEMS technology, the integration of the devices and the size of the chips continue to improve. It offers the following advantages: smaller device size, better electrical and mechanical properties, the possibility of mass production, integration in many areas, etc. [144,145]. These advantages are of great significance for the miniaturization, integration, and large-scale production of sensors. MEMS technology is also now deeply integrated with acceleration sensors. It is known that MEMS technology combined with optical inspection technology can realize the measurement of sub-angstrom-scale sensitivity and nanometer-scale acceleration, and the performance is close to the Brownian noise limit of mechanical structures. This sensitivity cannot be achieved by capacitive or piezoresistive methods, and increasing the mass of the acceleration sensor will be sensed at lower frequencies [146].

5.2.2. Process of MEMS Sensors

MEMS processes can be divided into the following categories: additive processes, subtractive processes, patterning, material property changes, and mechanical steps. Common additive processes include metal evaporation, metal sputtering, chemical vapor deposition of organics, chemical vapor deposition of inorganics, plasma-assisted chemical vapor deposition, thermal oxidation, and electroplating [147]. Subtractive processes include plasma etching, reactive ion etching, deep reactive ion etching, silicon wet chemical etching, and thin film wet chemical etching. Patterning mainly refers to photolithography processes. Material property changes include ion implantation, diffusion doping, and thermal annealing. Mechanical steps are polishing, wafer bonding, wafer slicing, lead bonding, and chip packaging [148].
Some commercial MEMS sensors are designed to be general-purpose with large bandwidths, and such sensors have large intrinsic frequencies; however, this can lead to poor sensitivity. To avoid this problem, dedicated MEMS sensors can be designed for different applications and different frequencies. For example, in civil structure health monitoring, low-seismic-level, low-frequency, and high-sensitivity acceleration signals are required, then MEMS accelerometers with low intrinsic frequency and high background noise can be designed to achieve the purpose [149]. Packaging technology, one of the key technologies for MEMS acceleration sensors, also has a critical impact on the performance of the overall device. Because of the influence of different package shells, die adhesives, and chip materials, thermal stresses are generated by temperature changes during packaging [150]. Li et al. found through theoretical analysis and simulation tests that the die adhesive dosage is inversely proportional to the sensitivity during packaging, and the binder thickness is inversely proportional to the residual stress after packaging. Uniform binder dosage and thickness with a suitable potting adhesive can ensure the consistency and improvement in device sensitivity [151].
Taking the piezoelectric MEMS sensor as an example, we introduce the flow of the additive process, subtractive process, and integrated process in detail. Firstly, these three processes are compared according to the conditions of device size, preparation speed, sensitivity, etc. (see Figure 10).
  • Addition process
Additive processes refer to the deposition of piezoelectric films on a silicon substrate with a suitable insulating layer, followed by micromachining of the surface or silicon block. The process starts with the sequential deposition of a passive layer, an active layer, and an electrode layer. Different thicknesses of SiO2, Si3N4, or polycrystalline silicon are first deposited on a substrate material such as Si to form the passive and sacrificial layers required for the sensor. In Figure 11a, the electrode films are deposited using sputtering and vacuum evaporation techniques, and the piezoelectric films are prepared using a combination of sputtering, chemical solution deposition, and evaporation techniques. To achieve the detection of charge and voltage in sensitive elements, the piezoelectric film is generally sandwiched between two electrode layers, or an interlocking electrode pattern is placed on top of the piezoelectric film. In the former case, the bottom electrode layer is used as a seed layer on which a textured piezoelectric film is deposited [152].
2.
Subtractive process
The subtractive process is the direct micromachining of single or polycrystalline piezoelectric materials, piezoelectric ceramics, etc., followed by appropriate plating [153] (see Figure 11b). It allows for the integration of sensors and actuators. It is worth mentioning that the etching of piezoelectric materials remains difficult for subtractive processes because of the stable nature of their oxides, which do not easily undergo the selective chemical reactions required for patterning. It is also the case that piezoelectric sensors implemented using batch micromachining processes lack a reliable and fast etching process. As a result, there are always limitations to such sensors [154].
3.
Integration process
Integration processes integrate micromechanical structures in silicon onto large piezoelectric substrates through bonding techniques, such as integrating precision micro-machined silicon structures onto PZT substrates. This has the advantage of being able to use a large number of Si micromachining techniques without the need to develop extensive micromachining capabilities for PZT. However, the success of this technique is predicated on the availability of low temperatures (preferably <200 °C) as well as precisely aligned bonding techniques. Andrea [154] et al. designed a MEMS actuator for PZT using an integrated process (see Figure 11c). This actuator is integrated with an embedded varistor, allowing the amplitude to be precisely controlled, and the stability and white noise to be effectively improved.
4.
Sensor fabrication on the silicon substrate
The basic flow of sensor fabrication on a silicon substrate is shown in Figure 11d [155], where spin-on glass (SOG) is first spin-coated onto silicon and dried at a certain temperature to cross-link the SOG. Next, an aluminum layer is evaporated to form an ohmic contact and pattern, and this is followed by alumina sputtering in an argon atmosphere (see Figure 11(d2)). The ZnO films are patterned using a mixture of a certain volume ratio of phosphoric acid, acetic acid, and aqueous solution. The second SOG layer is then spin-coated and patterned (see Figure 11(d3,d4)). The top electrode is then evaporated and patterned (see Figure 11(d5)). An SOG is spin-coated to encapsulate the sensor and is patterned, then the wafer is cut (see Figure 11(d6,d7)). Finally, it is processed using bulk micromachining or by cutting into strips (see Figure 11(d8)).
5.
Sensor fabrication on steel substrates
The manufacturing of sensors on steel substrates is more difficult compared to silicon [148], specifically related to two factors. One is that the suspension design of the steel substrate requires the steel sheet to be limited to a certain thickness and is therefore highly susceptible to damage during fabrication. After the vacuum chuck was invented to solve this problem, its use caused permanent deformation of the steel sheet, a deformation that, in turn, led to a change in the film thickness on the substrate. In response, the steel can be glued to the silicon shank with water and then the two wafers are heated to 100 °C, evaporating the water and releasing the steel substrate. The second is that steel has a large coefficient of thermal expansion. It is an order of magnitude greater than most micromachined compatible materials. This mismatch can lead to significant residual stresses and cause adhesion problems after film deposition.
For these two reasons, the temperature should be kept low during the manufacture of steel substrate sensors. First, the SOG is spin-coated onto the steel substrate and baked to flatten the steel substrate and prevent oxidation on the front side of the substrate. After the SOG layer has been applied to the steel substrate (Figure 11(e1)), the aluminum layer is evaporated and patterned to treat the bottom electrode (Figure 11(e2)). A second SOG layer is then coated; the benefit of using this additional SOG layer is its use as a buffer layer between the piezoelectric material and the aluminum layer, reducing residual stress gradients and the better adhesion of the piezoelectric material to the aluminum; it also allows the aluminum layer to be planarized for better deposition of zinc oxide. Continuing with the deposition and patterning of the ZnO layer on top of the second SOG layer, the SOG layer sandwiched between the ZnO and the underlying electrode is dry-etched in the SF plasma (Figure 11(e2)), a self-aligning process as the ZnO layer is used as a mask. The rest of the process is shown in Figure 11(e3–e7) and is the same as the process used in the Si substrate. After the sensor has been fabricated, the steel sheet is etched, as in Figure 11(e8), and finally tested for assembly.
Figure 11. Process implementation of MEMS piezoelectric acceleration sensors: additive process, subtractive process, integrated process, sensor manufacturing on silicon substrates, and sensor fabrication on steel substrates. Image for “integrated process”: Reprinted with permission from Ref. [154]. Copyright 2021, Elsevier.
Figure 11. Process implementation of MEMS piezoelectric acceleration sensors: additive process, subtractive process, integrated process, sensor manufacturing on silicon substrates, and sensor fabrication on steel substrates. Image for “integrated process”: Reprinted with permission from Ref. [154]. Copyright 2021, Elsevier.
Coatings 13 01252 g011

5.2.3. Application of MEMS Flexible Piezoelectric Accelerometer

Li et al. [156] proposed an LC wireless passive flexible acceleration sensor to eliminate the difficulty of measuring acceleration on the surface of curved structures. The principle of this accelerometer lies in varying the distance between the accelerometer and the antenna, using RF wireless transmission to measure the acceleration signal, and extending the lifetime by preventing the wire from falling off during vibration. It consists of a flexible polyimide (PI) substrate and a planar spiral inductor coil, and the processes for manufacturing such sensors are shown in Figure 12a. The measurements show that the acceleration sensor has a measurement range of 20 to 100 m/s2 and the acceleration and output voltage are almost linear with a maximum error of less than 0.037%. An ultrathin bendable flexible silicon-based piezoelectric sensor [157] was realized by wet anisotropic etching as a post-processing step, and the cutting of the film was achieved by a low-cost pre-etching cutting technique (see Figure 12b). Due to the optimized polarization procedure, this sensor has better piezoelectric properties and facilitates low-cost fabrication, as shown in Figure 12c. Figure 12d,e [158] show a flexible acceleration sensor composed of PVDF-TrFE and nanoclay composites that can be used for vibration measurement. The components of the sensor also include parallel stripe electrodes with impedance-matching circuits and an arc spring polyimide (PI) substrate. The test results in Figure 12f show that the design has high sensitivity, and the reasonable choice of the nanoclay composition can further promote the sensitivity. It is worth mentioning that this sensor also has good temperature stability and noise resistance.

6. Prospect and Outlook

In summary, the research progress of flexible piezoelectric acceleration sensors and their recent research results are introduced from the perspectives of properties, materials, structures, and processes. As a combination of flexible electronics, acceleration sensors, and other hot fields, flexible piezoelectric acceleration sensors make up for the shortcomings of traditional acceleration sensors, such as poor impact resistance and a complex manufacturing process. At the same time, it can achieve high flexibility, high sensitivity, self-power, and high reliability, especially suitable for acceleration measurement on some bending structures. Therefore, it has important applications in biomedicine, aerospace, wearable devices, the Internet of Things, and other fields.
Although the flexible piezoelectric acceleration sensor has made great progress and development in the past few years and has great potential and advantages in various fields at present, there are still some problems to be solved or improved in practical applications. Designing and fabricating high-performance piezoelectric sensors involve material innovation, structural innovation, material optimization, and principal innovation. However, most of the preparation processes are carried out under laboratory conditions, and due to the uncertainty of complexity, repeatability, and stability in practical applications, it is difficult to achieve the large-scale production of piezoelectric acceleration sensors under low-cost conditions. Three-dimensional printing technology and MEMS technology can be further improved to achieve the purpose of reducing costs to achieve mass production, which is a major aspect worth exploring. In addition, piezoelectric acceleration sensors with low power consumption meet the requirements of small size and low power consumption for future flexible microelectronic systems. Therefore, the development of integration technologies for flexible piezoelectric acceleration sensors and other electronic devices to achieve complete and complex functions is another hot research topic.

Author Contributions

Writing—original draft preparation, Y.L.; writing—review and editing, H.Y., Q.L., W.S., Y.C., X.C. and L.Q.; supervision, Q.L., X.C. and L.Q. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by The National Natural Science Foundation of China, grant number U2006218; The Project of Construction and Support for high-level Innovative Teams of Beijing Municipal Institutions, grant number BPHR20220124; and the Qin Xin Talents Cultivation Program, Beijing Information Science & Technology University, grant number QXTCP A202103.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Al-Turjman, F.; Nawaz, M.H.; Ulusar, U.D. Intelligence in the Internet of Medical Things era: A systematic review of current and future trends. Comput. Commun. 2020, 150, 644. [Google Scholar] [CrossRef]
  2. Yin, R.; Wang, D.; Zhao, S.; Lou, Z.; Shen, G. Wearable Sensors-Enabled Human–Machine Interaction Systems: From Design to Application. Adv. Funct. Mater. 2021, 31, 2008936. [Google Scholar] [CrossRef]
  3. Wang, Z.; Yang, Z.; Dong, T. A review of wearable technologies for elderly care that can accurately track indoor position, recognize physical activities and monitor vital signs in real time. Sensors 2017, 17, 341. [Google Scholar] [CrossRef] [Green Version]
  4. Abdulkarem, M.; Samsudin, K.; Rokhani, F.Z.; Rasid, M.F.A. Wireless sensor network for structural health monitoring: A contemporary review of technologies, challenges, and future direction. Struct. Health Monit. 2020, 19, 693. [Google Scholar] [CrossRef]
  5. Sunar, M.; Rao, S. Recent advances in sensing and control of flexible structures via piezoelectric materials technology. Appl. Mech. Rev. 1999, 52, 1–16. [Google Scholar] [CrossRef]
  6. Lu, Y. Industry 4.0: A survey on technologies, applications and open research issues. J. Ind. Inf. Integr. 2017, 6, 1–10. [Google Scholar] [CrossRef]
  7. Wu, J.; Guo, S.; Huang, H.; Liu, W.; Xiang, Y. Information and Communications Technologies for Sustainable Development Goals: State-of-the-Art, Needs and Perspectives. IEEE Commun. Surv. Tutor. 2018, 20, 2389. [Google Scholar] [CrossRef] [Green Version]
  8. Leal Filho, W.; Yang, P.; Eustachio, J.H.P.P.; Azul, A.M.; Gellers, J.C.; Gielczyk, A.; Dinis, M.A.P.; Kozlova, V. Deploying digitalisation and artificial intelligence in sustainable development research. Environ. Dev. Sustain. 2022, 25, 4957–4988. [Google Scholar] [CrossRef] [PubMed]
  9. Liao, Q.; Yin, Y.; Zhang, J.; Si, W.; Hou, W.; Qin, L. In Situ Growth of Nanosilver on Fabric for Flexible Stretchable Electrodes. Int. J. Mol. Sci. 2022, 23, 13236. [Google Scholar] [CrossRef] [PubMed]
  10. Hou, W.; Liao, Q.; Xie, S.; Song, Y.; Qin, L. Prospects and Challenges of Flexible Stretchable Electrodes for Electronics. Coatings 2022, 12, 558. [Google Scholar] [CrossRef]
  11. Mohankumar, P.; Ajayan, J.; Yasodharan, R.; Devendran, P.; Sambasivam, R. A review of micromachined sensors for automotive applications. Measurement 2019, 140, 305. [Google Scholar] [CrossRef]
  12. Ai, C.; Zhao, X.; Wen, D. Characteristics Research of a High Sensitivity Piezoelectric MOSFET Acceleration Sensor. Sensors 2020, 20, 4988. [Google Scholar] [CrossRef] [PubMed]
  13. Deng, W.; Yang, T.; Jin, L.; Yan, C.; Huang, H.; Chu, X.; Wang, Z.; Xiong, D.; Tian, G.; Gao, Y.; et al. Cowpea-structured PVDF/ZnO nanofibers based flexible self-powered piezoelectric bending motion sensor towards remote control of gestures. Nano Energy 2019, 55, 516. [Google Scholar] [CrossRef]
  14. Cui, H.; Hensleigh, R.; Yao, D.; Maurya, D.; Kumar, P.; Kang, M.G.; Priya, S.; Zheng, X. Three-dimensional printing of piezoelectric materials with designed anisotropy and directional response. Nat. Mater. 2019, 18, 234. [Google Scholar] [CrossRef]
  15. Pan, M.; Yuan, C.; Liang, X.; Zou, J.; Zhang, Y.; Bowen, C. Triboelectric and Piezoelectric Nanogenerators for Future Soft Robots and Machines. iScience 2020, 23, 101682. [Google Scholar] [CrossRef] [PubMed]
  16. Navaraj, W.; Dahiya, R. Fingerprint-enhanced capacitive-piezoelectric flexible sensing skin to discriminate static and dynamic tactile stimuli. Adv. Intell. Syst. 2019, 1, 1900051. [Google Scholar] [CrossRef]
  17. Park, Y.; Shim, J.; Jeong, S.; Yi, G.-R.; Chae, H.; Bae, J.W.; Kim, S.O.; Pang, C. Microtopography-Guided Conductive Patterns of Liquid-Driven Graphene Nanoplatelet Networks for Stretchable and Skin-Conformal Sensor Array. Adv. Mater. 2017, 29, 1606453. [Google Scholar] [CrossRef]
  18. Cao, R.; Pu, X.; Du, X.; Yang, W.; Wang, J.; Guo, H.; Zhao, S.; Yuan, Z.; Zhang, C.; Li, C. Screen-printed washable electronic textiles as self-powered touch/gesture tribo-sensors for intelligent human–machine interaction. ACS Nano 2018, 12, 5190. [Google Scholar] [CrossRef]
  19. Lowe, R.; Andreasson, R.; Alenljung, B.; Lund, A.; Billing, E. Designing for a Wearable Affective Interface for the NAO Robot: A Study of Emotion Conveyance by Touch. Multimodal Technol. Interact. 2018, 2, 2. [Google Scholar] [CrossRef] [Green Version]
  20. Han, J.H.; Min, B.K.; Hong, S.K.; Park, H.; Kwak, J.-H.; Wang, H.S.; Joe, D.J.; Park, J.H.; Jung, Y.H.; Hur, S.; et al. Machine learning-based self-powered acoustic sensor for speaker recognition. Nano Energy 2018, 53, 658–665. [Google Scholar] [CrossRef]
  21. Jin, L.; Ma, S.; Deng, W.; Yan, C.; Yang, T.; Chu, X.; Tian, G.; Xiong, D.; Lu, J.; Yang, W. Polarization-free high-crystallization β-PVDF piezoelectric nanogenerator toward self-powered 3D acceleration sensor. Nano Energy 2018, 50, 632. [Google Scholar] [CrossRef]
  22. da Fonseca, I.M.; Rade, D.A.; Goes, L.C.S.; de Paula Sales, T. Attitude and vibration control of a satellite containing flexible solar arrays by using reaction wheels, and piezoelectric transducers as sensors and actuators. Acta Astronaut. 2017, 139, 357. [Google Scholar] [CrossRef]
  23. Chen, X.; Zeng, Q.; Shao, J.; Li, S.; Luo, Y. Channel-Crack-Designed Suspended Sensing Membrane as a Fully Flexible Vibration Sensor with High Sensitivity and Dynamic Range. ACS Appl. Mater. Interfaces 2021, 13, 34637–34647. [Google Scholar] [CrossRef] [PubMed]
  24. Yuan, X.; Gao, X.; Shen, X.; Yang, J.; Dong, S. A 3D-printed, alternatively tilt-polarized PVDF-TrFE polymer with enhanced piezoelectric effect for self-powered sensor application. Nano Energy 2021, 85, 105985. [Google Scholar] [CrossRef]
  25. Smart Cities. Available online: https://www.freepik.com/free-vector/isometric-smart-city-background_4358963.htm (accessed on 27 February 2023).
  26. Smart Farming. Available online: https://www.freepik.com/free-photo/smart-robotic-farmers-concept-robot-farmers-agriculture-technology-farm-automation_21544365.htm (accessed on 27 February 2023).
  27. Smart Buildings. Available online: https://www.freepik.com/free-vector/isometric-smart-city-background_4358967.htm (accessed on 4 November 2022).
  28. Armin, D.; Jakob, R.; Gregor, E.; Miha, B. On the performance of direct piezoelectric rotational accelerometers in experimental structural dynamics. Measurement 2018, 127, 292. [Google Scholar] [CrossRef]
  29. Chen, X.; Luo, L.; Zeng, Z.; Jiao, J.; Shehzad, M.; Yuan, G.; Luo, H.; Wang, Y. Bio-inspired flexible vibration visualization sensor based on piezo-electrochromic effect. J. Mater. 2020, 6, 643–650. [Google Scholar] [CrossRef]
  30. Zheng, L.; Shuai, C.; Wang, L.; Shi, R.; Shen, G. Ultrasensitive and ultraflexible e-skins with dual functionalities for wearable electronics. Nano Energy 2017, 38, 28–35. [Google Scholar] [CrossRef]
  31. Tsukamoto, T.; Umino, Y.; Shiomi, S.; Yamada, K.; Suzuki, T. Bimorph piezoelectric vibration energy harvester with flexible 3D meshed-core structure for low frequency vibration. Sci. Technol. Adv. Mater. 2018, 19, 660. [Google Scholar] [CrossRef] [Green Version]
  32. Yamamoto, Y.; Harada, S.; Yamamoto, D.; Honda, W.; Arie, T.; Akita, S.; Takei, K. Printed multifunctional flexible device with an integrated motion sensor for health care monitoring. Sci. Adv. 2016, 2, e1601473. [Google Scholar] [CrossRef] [Green Version]
  33. Pang, C.; Koo, J.H.; Nguyen, A.; Caves, J.M.; Kim, M.G.; Chortos, A.; Kim, K.; Wang, P.J.; Tok, J.B.H.; Bao, Z. Highly Skin-Conformal Microhairy Sensor for Pulse Signal Amplification. Adv. Mater. 2015, 27, 634–640. [Google Scholar] [CrossRef]
  34. Wu, Y.; Ma, Y.; Zheng, H.; Ramakrishna, S. Piezoelectric materials for flexible and wearable electronics: A review. Mater. Des. 2021, 211, 110164. [Google Scholar] [CrossRef]
  35. Chen, X.; Li, X.; Shao, J.; An, N.; Tian, H.; Wang, C.; Han, T.; Wang, L.; Lu, B. High-Performance Piezoelectric Nanogenerators with Imprinted P(VDF-TrFE)/BaTiO3 Nanocomposite Micropillars for Self-Powered Flexible Sensors. Small 2017, 13, 1604245. [Google Scholar] [CrossRef] [PubMed]
  36. Huang, Y.; Fan, X.; Chen, S.C.; Zhao, N. Emerging Technologies of Flexible Pressure Sensors: Materials, Modeling, Devices, and Manufacturing. Adv. Funct. Mater. 2019, 29, 1808509. [Google Scholar] [CrossRef]
  37. Bergmayr, T.; Kralovec, C.; Schagerl, M. Vibration-Based Thermal Health Monitoring for Face Layer Debonding Detection in Aerospace Sandwich Structures. Appl. Sci. 2020, 11, 211. [Google Scholar] [CrossRef]
  38. Zhang, S.; Chen, Z.; Li, G.; Wu, Y. A novel Z-shaped elastic flange structure for increasing the amplitude output of a piezoelectric ultrasonic transducer. Sens. Actuators A Phys. 2021, 331, 112995. [Google Scholar] [CrossRef]
  39. Pan, Q.S.; He, L.G.; Pan, C.L.; Xiao, G.J.; Feng, Z.H. Resonant-type inertia linear motor based on the harmonic vibration synthesis of piezoelectric bending actuator. Sens. Actuators A Phys. 2014, 209, 169. [Google Scholar] [CrossRef]
  40. Levinzon, F.A. Noise of piezoelectric accelerometer with integral FET amplifier. IEEE Sens. J. 2005, 5, 1235. [Google Scholar] [CrossRef]
  41. Gesing, A.L.; Alves, F.D.P.; Paul, S.; Cordioli, J.A. On the design of a MEMS piezoelectric accelerometer coupled to the middle ear as an implantable sensor for hearing devices. Sci. Rep. 2018, 8, 3920. [Google Scholar] [CrossRef] [Green Version]
  42. Li, W.; Weldon, J.A.; Huang, Y.; Wang, K. Design and Simulation of Piezoelectric-Charge-Gated Thin-Film Transistor for Tactile Sensing. IEEE Electron Device Lett. 2016, 37, 325. [Google Scholar] [CrossRef]
  43. Guo, Z.Q.; Liu, L.Y.; Liu, J.; Wu, N.J. Noise analysis for high speed CMOS image sensor. Sel. Pap. Conf. Photoelectron. Technol. Comm. Chin. Soc. Astronaut. 2015, 9522, 182–187. [Google Scholar] [CrossRef]
  44. Shen, Z.; Tan, C.Y.; Yao, K.; Zhang, L.; Chen, Y.F. A miniaturized wireless accelerometer with micromachined piezoelectric sensing element. Sens. Actuators A Phys. 2016, 241, 113. [Google Scholar] [CrossRef]
  45. Kollias, A.T.; Avaritsiotis, J.N. A study on the performance of bending mode piezoelectric accelerometers. Sens. Actuators A Phys. 2005, 121, 434. [Google Scholar] [CrossRef]
  46. Feng, G.H.; Pan, Y.L. Establishing a cost-effective sensing system and signal processing method to diagnose preload levels of ball screws. Mech. Syst. Signal Process. 2012, 28, 78. [Google Scholar] [CrossRef]
  47. Jiang, C.; Liu, X.; Yu, F.P.; Zhang, S.; Zhao, X. High-temperature vibration sensor based on Ba2TiSi2O8 piezoelectric crystal with ultra-stable sensing performance up to 650 °C. IEEE Trans. Ind. Electron. 2020, 68, 12850–12859. [Google Scholar] [CrossRef]
  48. Zhang, S.; Jiang, X.; Lapsley, M.; Moses, P.; Shrout, T.R. Piezoelectric Accelerometers for Ultrahigh Temperature Application. Appl. Phys. Lett. 2010, 96, 013506. [Google Scholar] [CrossRef]
  49. Bochobza-Degani, O.; Dan, J.S.; Socher, E.; Nemirovsky, Y. Design and noise consideration of an accelerometer employing modulated integrative differential optical sensing. Sens. Actuators A Phys. 2000, 84, 53. [Google Scholar] [CrossRef]
  50. Yaghootkar, B.; Azimi, S.; Bahreyni, B. A High-Performance Piezoelectric Vibration Sensor. IEEE Sens. J. 2017, 17, 4005. [Google Scholar] [CrossRef]
  51. Yan, L.; Zhao, Y.; Wang, W.; Lu, S.; Jiang, Z. A high-performance multi-beam microaccelerometer for vibration monitoring in intelligent manufacturing equipment. Sens. Actuators A Phys. 2013, 189, 8–16. [Google Scholar] [CrossRef]
  52. Roy, A.L.; Sarkar, H.; Dutta, A.; Bhattacharyya, T.K. A high precision SOI MEMS–CMOS ±4 g piezoresistive accelerometer. Sens. Actuators A Phys. 2014, 210, 77. [Google Scholar] [CrossRef]
  53. Speller, K.E.; Yu, D. A low-noise MEMS accelerometer for unattended ground sensor applications. In Unattended/Unmanned Ground, Ocean, and Air Sensor Technologies and Applications; SPIE: Bellingham, WA, USA, 2004; Volume 5417, pp. 63–72. [Google Scholar] [CrossRef]
  54. Wang, L.P.; Deng, K.; Zou, L.; Wolf, R.; Davis, R.J.; Trolier-Mckinstry, S. Microelectromechanical systems (MEMS) accelerometers using lead zirconate titanate thick films. Electron Device Lett. IEEE 2002, 23, 182. [Google Scholar] [CrossRef]
  55. Gerfers, F.; Kohlstadt, M.; Bar, H.; He, M.Y.; Wang, L.P. Sub-μg ultra-low-noise MEMS accelerometers based on CMOS-compatible piezoelectric AlN thin films. In Proceedings of the TRANSDUCERS 2007-2007 International Solid-State Sensors, Actuators and Microsystems Conference, Lyon, France, 10–14 June 2007; pp. 1191–1194. [Google Scholar] [CrossRef]
  56. PCB Piezotronics MTS Systems Corp. Available online: https://www.pcb.com/products?m=357B69 (accessed on 21 February 2023).
  57. Shi, J.; Liu, S.; Zhang, L.; Yang, B.; Shu, L.; Yang, Y.; Ren, M.; Wang, Y.; Chen, J.; Chen, W. Smart Textile-Integrated Microelectronic Systems for Wearable Applications. Adv. Mater. 2020, 32, 1901958. [Google Scholar] [CrossRef] [PubMed]
  58. Liao, Q.; Hou, W.; Liao, K.; Chen, L.; Song, Y.; Gao, G.; Qin, L. Solid-phase sintering and vapor-liquid-solid growth of BP@MgO quantum dot crystals with a high piezoelectric response. J. Adv. Ceram. 2022, 11, 1725. [Google Scholar] [CrossRef]
  59. Hofmann, P.; Walch, A.; Dinkelmann, A.; Selvarayan, S.K.; Gresser, G.T. Woven piezoelectric sensors as part of the textile reinforcement of fiber reinforced plastics. Compos. Part A Appl. Sci. Manuf. 2019, 116, 79. [Google Scholar] [CrossRef]
  60. Mahdavi, Z.; Khodakarami, M.I.; Kontoni, D.-P. Qualitative flaw detection in piezoelectric material for elastostatics using linear sampling method. In Proceedings of the 6th IC-EpsMsO, Athens, Greece, 8–11 July 2015. [Google Scholar]
  61. Stoynov, Y.; Rangelov, T. Inverse Scattering Problems for Cracked Piezoelectric Media. Aip Conf. Am. Inst. Phys. 2013, 1561, 318–326. [Google Scholar] [CrossRef]
  62. Valagiannopoulos, C. On measuring the permittivity tensor of an anisotropic material from the transmission coefficients. Prog. Electromagn. Res. B 2008, 9, 105. [Google Scholar] [CrossRef] [Green Version]
  63. Valagiannopoulos, C.A.; Tsitsas, N.L. Integral equation analysis of a low-profile receiving planar microstrip antenna with a cloaking superstrate. Radio Sci. 2012, 47, 1–12. [Google Scholar] [CrossRef]
  64. Boryło, P.; Matus, K.; Lukaszkowicz, K.; Kubacki, J.; Balin, K.; Basiaga, M.; Szindler, M.; Mikuła, J. The influence of atomic layer deposition process temperature on ZnO thin film structure. Appl. Surf. Sci. 2019, 474, 177. [Google Scholar] [CrossRef]
  65. Han, X.G.; He, H.Z.; Kuang, Q.; Zhou, X.; Zhang, X.H.; Xu, T.; Xie, Z.X.; Zheng, L.S. Controlling Morphologies and Tuning the Related Properties of Nano/Microstructured ZnO Crystallites. J. Phys. Chem. C 2009, 113, 584. [Google Scholar] [CrossRef]
  66. Qian, Y.K.; Kang, D.J. Poly(dimethylsiloxane)/ZnO Nanoflakes/Three-Dimensional Graphene Heterostructures for High-Performance Flexible Energy Harvesters with Simultaneous Piezoelectric and Triboelectric Generation. ACS Appl. Mater. Interfaces 2018, 10, 32281–32288. [Google Scholar] [CrossRef]
  67. Jin, W.; Durussel, A.; Sandu, C.S.; Sahini, M.G.; He, Z.; Setter, N. Mechanism of hydrothermal growth of ferroelectric PZT nanowires. J. Cryst. Growth 2012, 347, 1–6. [Google Scholar] [CrossRef]
  68. Shin, D.-J.; Jeong, S.-J.; Seo, C.-E.; Cho, K.-H.; Koh, J.-H. Multi-layered piezoelectric energy harvesters based on PZT ceramic actuators. Ceram. Int. 2015, 41, S686. [Google Scholar] [CrossRef]
  69. Tian, G.; Deng, W.; Gao, Y.; Xiong, D.; Yan, C.; He, X.; Yang, T.; Jin, L.; Chu, X.; Zhang, H.; et al. Rich lamellar crystal baklava-structured PZT/PVDF piezoelectric sensor toward individual table tennis training. Nano Energy 2019, 59, 574. [Google Scholar] [CrossRef]
  70. Sood, A.; Desseigne, M.; Dev, A.; Maurizi, L.; Kumar, A.; Millot, N.; Han, S.S. A Comprehensive Review on Barium Titanate Nanoparticles as a Persuasive Piezoelectric Material for Biomedical Applications: Prospects and Challenges. Small 2023, 19, 2206401. [Google Scholar] [CrossRef]
  71. Upadhyay, S.; Shrivastava, J.; Solanki, A.; Choudhary, S.; Sharma, V.; Kumar, P.; Singh, N.; Satsangi, V.R.; Shrivastav, R.; Waghmare, U.V.; et al. Enhanced Photoelectrochemical Response of BaTiO3 with Fe Doping: Experiments and First-Principles Analysis. J. Phys. Chem. C 2011, 115, 24373–24380. [Google Scholar] [CrossRef]
  72. Selvarajan, S.; Alluri, N.R.; Chandrasekhar, A.; Kim, S.J. Unconventional active biosensor made of piezoelectric BaTiO3 nanoparticles for biomolecule detection. Sens. Actuators B Chem. 2017, 253, 1180–1187. [Google Scholar] [CrossRef]
  73. Sharma, T.; Je, S.S.; Gill, B.; Zhang, J. Patterning piezoelectric thin film PVDF–TrFE based pressure sensor for catheter application. Sens. Actuators A Phys. 2012, 177, 87. [Google Scholar] [CrossRef]
  74. Rajabi, H.; Ghaemi, N.; Madaeni, S.S.; Daraei, P.; Khadivi, M.A.; Falsafi, M. Nanoclay embedded mixed matrix PVDF nanocomposite membrane: Preparation, characterization and biofouling resistance. Appl. Surf. Sci. 2014, 313, 207. [Google Scholar] [CrossRef]
  75. Kalimuldina, G.; Turdakyn, N.; Abay, I.; Medeubayev, A.; Nurpeissova, A.; Adair, D.; Bakenov, Z. A Review of Piezoelectric PVDF Film by Electrospinning and Its Applications. Sensors 2020, 20, 5214. [Google Scholar] [CrossRef]
  76. Duan, S.; Wu, J.; Xia, J.; Lei, W. Innovation Strategy Selection Facilitates High-Performance Flexible Piezoelectric Sensors. Sensors 2020, 20, 2820. [Google Scholar] [CrossRef]
  77. Ghosh, S.K.; Mandal, D. Synergistically enhanced piezoelectric output in highly aligned 1D polymer nanofibers integrated all-fiber nanogenerator for wearable nano-tactile sensor. Nano Energy 2018, 53, 245–257. [Google Scholar] [CrossRef]
  78. Meng, X.; Zhang, Z.; Lin, D.; Liu, W.; Zhou, S.; Ge, S.; Su, Y.; Peng, C.; Zhang, L. Effects of particle size of dielectric fillers on the output performance of piezoelectric and triboelectric nanogenerators. J. Adv. Ceram. 2021, 10, 991. [Google Scholar] [CrossRef]
  79. Qi, Y.; Pan, L.; Ma, L.; Liao, P.; Ge, J.; Zhang, D.; Zheng, Q.; Yu, B.; Tang, Y.; Sun, D. Investigation on FT-IR spectra and dielectric property of PVDF/inorganic composites. J. Mater. Sci. Mater. Electron. 2013, 24, 1446. [Google Scholar] [CrossRef]
  80. Newnham, R.E.; Skinner, D.P.; Cross, L.E. Connectivity and Piezoelectric–Pyroelectric Composite. Mater. Res. Bull. 1978, 13, 525. [Google Scholar] [CrossRef]
  81. Garg, T.; Annapureddy, V.; Sekhar, K.C.; Jeong, D.Y.; Dabra, N.; Hundal, J.S. Modulation in polymer properties in PVDF/BCZT composites with ceramic content and their energy density capabilities. Polym. Compos. 2020, 41, 5305–5316. [Google Scholar] [CrossRef]
  82. Zhen, L.; Lu, L.; Yao, Y.; Liu, J.; Yang, B. Flexible inorganic piezoelectric functional films and their applications. J. Adv. Ceram. 2023, 12, 433. [Google Scholar] [CrossRef]
  83. Mao, Y.; Zhao, P.; Mcconohy, G.; Yang, H.; Tong, Y.; Wang, X. Sponge-Like Piezoelectric Polymer Films for Scalable and Integratable Nanogenerators and Self-Powered Electronic Systems. Adv. Energy Mater. 2014, 4, 1301624. [Google Scholar] [CrossRef]
  84. Yang, T.; Pan, H.; Tian, G.; Zhang, B.; Yang, W. Hierarchically Structured PVDF/ZnO Core-shell Nanofibers for Self-powered Physiological Monitoring Electronics. Nano Energy 2020, 72, 104706. [Google Scholar] [CrossRef]
  85. Arshad, A.N.; Wahid, M.; Rusop, M.; Majid, W.; Subban, R.; Rozana, M. Dielectric and Structural Properties of Poly(vinylidene fluoride) (PVDF) and Poly(vinylidene fluoride-trifluoroethylene) (PVDF-TrFE) Filled with Magnesium Oxide Nanofillers. J. Nanomater. 2019, 2019, 5961563. [Google Scholar] [CrossRef]
  86. Alluri, N.R.; Chandrasekhar, A.; Vivekananthan, V.; Purusothaman, Y.; Selvarajan, S.; Jeong, J.H.; Kim, S.-J. Scavenging Biomechanical Energy Using High-Performance, Flexible BaTiO3 Nanocube/PDMS Composite Films. ACS Sustain. Chem. Eng. 2017, 5, 4730. [Google Scholar] [CrossRef]
  87. Liao, Q.; Hou, W.; Zhang, J.; Qin, L. Controllable Preparation of Silver Nanowires and Its Application in Flexible Stretchable Electrode. Coatings 2022, 12, 1756. [Google Scholar] [CrossRef]
  88. Wolf, M.P.; Salieb-Beugelaar, G.B.; Patrick, H. PDMS with designer functionalities—Properties, modifications strategies, and applications. Prog. Polym. Sci. 2018, 83, 97. [Google Scholar] [CrossRef]
  89. Li, Z.; Peng, Y.; Xu, Z.; Peng, J.; Pu, H. Harnessing energy from suspension systems of oceanic vehicles with high-performance piezoelectric generators. Energy 2021, 228, 120523. [Google Scholar] [CrossRef]
  90. Chen, Y.; Xue, X. Advances in the Structural Health Monitoring of Bridges Using Piezoelectric Transducers. Sensors 2018, 18, 4312. [Google Scholar] [CrossRef] [Green Version]
  91. Jiang, X.; Kim, K.; Zhang, S.; Johnson, J.; Salazar, G. High-Temperature Piezoelectric Sensing. Sensors 2014, 14, 144. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Wang, Y.; Song, P.; Li, X.; Ru, C.; Ferrari, G.; Balasubramanian, P.; Amabili, M.; Sun, Y.; Liu, X. A Paper-Based Piezoelectric Accelerometer. Micromachines 2018, 9, 19. [Google Scholar] [CrossRef] [Green Version]
  93. Jeong, C.K.; Baek, C.; Kingon, A.I.; Park, K.-I.; Kim, S.-H. Lead-Free Perovskite Nanowire-Employed Piezopolymer for Highly Efficient Flexible Nanocomposite Energy Harvester. Small 2018, 14, 1704022. [Google Scholar] [CrossRef]
  94. Yuan, Y.; Chen, H.; Xu, H.; Jin, Y.; Chen, G.; Zheng, W.; Wang, W.; Wang, Y.; Gao, L. Highly sensitive and wearable bionic piezoelectric sensor for human respiratory monitoring. Sens. Actuators A Phys. 2022, 345, 113818. [Google Scholar] [CrossRef]
  95. Zhang, B.; Zhang, L.; Deng, W.; Jin, L.; Chun, F.; Pan, H.; Gu, B.; Zhang, H.; Lv, Z.; Yang, W.; et al. Self-Powered Acceleration Sensor Based on Liquid Metal Triboelectric Nanogenerator for Vibration Monitoring. ACS Nano 2017, 11, 7440. [Google Scholar] [CrossRef]
  96. Chen, J.; Oh, S.K.; Nabulsi, N.; Johnson, H.; Wang, W.; Ryou, J.-H. Biocompatible and sustainable power supply for self-powered wearable and implantable electronics using III-nitride thin-film-based flexible piezoelectric generator. Nano Energy 2019, 57, 670. [Google Scholar] [CrossRef]
  97. Chiu, Y.; Lin, W.; Wang, H.; Huang, S.; Wu, M. Development of a piezoelectric polyvinylidene fluoride (PVDF) polymer-based sensor patch for simultaneous heartbeat and respiration monitoring. Sens. Actuators A Phys. 2013, 189, 328. [Google Scholar] [CrossRef]
  98. MacDonald, E.; Wicker, R. Multiprocess 3D printing for increasing component functionality. Science 2016, 353, aaf2093. [Google Scholar] [CrossRef] [PubMed]
  99. Kalkal, A.; Kumar, S.; Kumar, P.; Pradhan, R.; Willander, M.; Packirisamy, G.; Kumar, S.; Malhotra, B.D. Recent advances in 3D printing technologies for wearable (bio)sensors. Addit. Manuf. 2021, 46, 102088. [Google Scholar] [CrossRef]
  100. Park, S.; Shou, W.; Makatura, L.; Matusik, W.; Fu, K. 3D printing of polymer composites: Materials, processes, and applications. Matter 2022, 5, 43. [Google Scholar] [CrossRef]
  101. Xu, Y.; Wu, X.; Guo, X.; Kong, B.; Zhang, M.; Qian, X.; Mi, S.; Sun, W. The Boom in 3D-Printed Sensor Technology. Sensors 2017, 17, 1166. [Google Scholar] [CrossRef] [PubMed]
  102. Leigh, S.J.; Bradley, R.J.; Purssell, C.P.; Billson, D.R.; Hutchins, D.A. A Simple, Low-Cost Conductive Composite Material for 3D Printing of Electronic Sensors. PLoS ONE 2012, 7, e49365. [Google Scholar] [CrossRef]
  103. Yao, D.; Cui, H.; Hensleigh, R.; Smith, P.; Alford, S.; Bernero, D.; Bush, S.; Mann, K.; Wu, H.F.; Chin-Nieh, M.; et al. Achieving the Upper Bound of Piezoelectric Response in Tunable, Wearable 3D Printed Nanocomposites. Adv. Funct. Mater. 2019, 29, 1903866. [Google Scholar] [CrossRef]
  104. Bodkhe, S.; Turcot, G.; Gosselin, F.P.; Therriault, D. One-Step Solvent Evaporation-Assisted 3D Printing of Piezoelectric PVDF Nanocomposite Structures. ACS Appl. Mater. Interfaces 2017, 9, 20833. [Google Scholar] [CrossRef]
  105. Hull, C.W. Apparatus for Production of Three-Dimensional Objects by Stereolithography. U.S. Patent 638905, 8 August 1984. [Google Scholar]
  106. Zhang, F.; Wei, M.; Viswanathan, V.V.; Swart, B.; Shao, Y.; Wu, G.; Zhou, C. 3D printing technologies for electrochemical energy storage. Nano Energy 2017, 40, 418. [Google Scholar] [CrossRef]
  107. Kim, H.; Torres, F.; Wu, Y.; Villagran, D.; Lin, Y.; Tseng, T.-L. Integrated 3D printing and corona poling process of PVDF piezoelectric films for pressure sensor application. Smart Mater. Struct. 2017, 26, 085027. [Google Scholar] [CrossRef]
  108. Kim, H.; Torres, F.; Islam, M.T.; Islam, M.D.; Chavez, L.A.; Garcia Rosales, C.A.; Wilburn, B.R.; Stewart, C.M.; Noveron, J.C.; Tseng, T.-L.B.; et al. Increased piezoelectric response in functional nanocomposites through multiwall carbon nanotube interface and fused-deposition modeling three-dimensional printing. MRS Commun. 2017, 7, 960. [Google Scholar] [CrossRef] [Green Version]
  109. Muñoz, J.; Pumera, M. 3D-printed biosensors for electrochemical and optical applications. TrAC Trends Anal. Chem. 2020, 128, 115933. [Google Scholar] [CrossRef]
  110. Renteria, A.; Balcorta, V.H.; Marquez, C.; Rodriguez, A.A.; Renteria-Marquez, I.; Regis, J.; Wilburn, B.; Patterson, S.; Espalin, D.; Tseng, T.-L.; et al. Direct ink write multi-material printing of PDMS-BTO composites with MWCNT electrodes for flexible force sensors. Flex. Print. Electron. 2022, 7, 015001. [Google Scholar] [CrossRef]
  111. Bogue, R. 3D printing: The dawn of a new era in manufacturing? Assem. Autom. 2013, 33, 307. [Google Scholar] [CrossRef]
  112. Gross, B.C.; Erkal, J.L.; Lockwood, S.Y.; Chen, C.; Spence, D.M. Evaluation of 3D Printing and Its Potential Impact on Biotechnology and the Chemical Sciences. Anal. Chem. 2014, 86, 3240. [Google Scholar] [CrossRef]
  113. Tiller, B.; Reid, A.; Zhu, B.; Guerreiro, J.; Domingo-Roca, R.; Curt Jackson, J.; Windmill, J.F.C. Piezoelectric microphone via a digital light processing 3D printing process. Mater. Des. 2019, 165, 107593. [Google Scholar] [CrossRef]
  114. He, Y.; Wu, Y.; Fu, J.; Gao, Q.; Qiu, J. Developments of 3D Printing Microfluidics and Applications in Chemistry and Biology: A Review. Electroanalysis 2016, 28, 1658. [Google Scholar] [CrossRef]
  115. Fang, C.; Fang, Z. Application of 3D Printing Technology in Hepato-Biliary-Pancreatic Surgery. In Biliary Tract Surgery; Fang, C., Lau, W.Y., Eds.; Springer: Singapore, 2021; pp. 117–130. [Google Scholar] [CrossRef]
  116. Dutta, D.; Prinz, F.B.; Rosen, D.; Weiss, L. Layered Manufacturing: Current Status and Future Trends. J. Comput. Inf. Sci. Eng. 2000, 1, 60. [Google Scholar] [CrossRef] [Green Version]
  117. Ko, S.H.; Pan, H.; Grigoropoulos, C.P.; Luscombe, C.K.; Fréchet, J.M.J.; Poulikakos, D. All-inkjet-printed flexible electronics fabrication on a polymer substrate by low-temperature high-resolution selective laser sintering of metal nanoparticles. Nanotechnology 2007, 18, 345202. [Google Scholar] [CrossRef]
  118. Nadgorny, M.; Ameli, A. Functional Polymers and Nanocomposites for 3D Printing of Smart Structures and Devices. ACS Appl. Mater. Interfaces 2018, 10, 17489–17507. [Google Scholar] [CrossRef]
  119. Song, S.; Li, Y.; Wang, Q.; Zhang, C. Boosting piezoelectric performance with a new selective laser sintering 3D printable PVDF/graphene nanocomposite. Compos. Part A Appl. Sci. Manuf. 2021, 147, 106452. [Google Scholar] [CrossRef]
  120. Pugalendhi, A.; Ranganathan, R. A review of additive manufacturing applications in ophthalmology. Proc. Inst. Mech. Eng. Part H J. Eng. Med. 2021, 235, 1146. [Google Scholar] [CrossRef] [PubMed]
  121. Zhao, W.; Wang, Z.; Zhang, J.; Wang, X.; Xu, Y.; Ding, N.; Peng, Z. Vat Photopolymerization 3D Printing of Advanced Soft Sensors and Actuators: From Architecture to Function. Adv. Mater. Technol. 2021, 6, 2001218. [Google Scholar] [CrossRef]
  122. Nightingale, A.M.; Leong, C.L.; Burnish, R.A.; Hassan, S.-u.; Zhang, Y.; Clough, G.F.; Boutelle, M.G.; Voegeli, D.; Niu, X. Monitoring biomolecule concentrations in tissue using a wearable droplet microfluidic-based sensor. Nat. Commun. 2019, 10, 2741. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Changyong, L.; Ninggui, H.; Feng, X.; Junda, T.; Zhangwei, C.; Xuchun, G.; Yuelong, F.; Changshi, L. 3D Printing Technologies for Flexible Tactile Sensors toward Wearable Electronics and Electronic Skin. Polymers 2018, 10, 629. [Google Scholar] [CrossRef] [Green Version]
  124. Turner, B.N.; Strong, R.; Gold, S.A. A review of melt extrusion additive manufacturing processes: I. Process design and modeling. Rapid Prototyp. J. 2014, 20, 192. [Google Scholar] [CrossRef]
  125. Abshirini, M.; Charara, M.; Marashizadeh, P.; Saha, M.C.; Altan, M.C.; Liu, Y. Functional nanocomposites for 3D printing of stretchable and wearable sensors. Appl. Nanosci. 2019, 9, 2071. [Google Scholar] [CrossRef]
  126. Davoodi, E.; Fayazfar, H.; Liravi, F.; Jabari, E.; Toyserkani, E. Drop-on-demand high-speed 3D printing of flexible milled carbon fiber/silicone composite sensors for wearable biomonitoring devices. Addit. Manuf. 2020, 32, 101016. [Google Scholar] [CrossRef]
  127. Barši Palmić, T.; Slavič, J.; Boltežar, M. Process Parameters for FFF 3D-Printed Conductors for Applications in Sensors. Sensors 2020, 20, 4542. [Google Scholar] [CrossRef]
  128. Spivey, E.C.; Xhemalçe, B.; Shear, J.B.; Finkelstein, I.J. 3D-printed microfluidic microdissector for high-throughput studies of cellular aging. Anal. Chem. 2014, 86, 7406. [Google Scholar] [CrossRef] [Green Version]
  129. Au, A.K.; Lee, W.; Folch, A. Mail-order microfluidics: Evaluation of stereolithography for the production of microfluidic devices. Lab A Chip 2014, 14, 1294. [Google Scholar] [CrossRef] [Green Version]
  130. Ni, Y.; Ji, R.; Long, K.; Bu, T.; Chen, K.; Zhuang, S. A review of 3D-printed sensors. Appl. Spectrosc. Rev. 2017, 52, 623. [Google Scholar] [CrossRef]
  131. Shamsinejad, S.; Flaviis, F.D.; Mousavi, P. Microstrip-Fed 3-D Folded Slot Antenna on Cubic Structure. IEEE Antennas Wirel. Propag. Lett. 2016, 15, 1081. [Google Scholar] [CrossRef]
  132. Adams, J.J.; Slimmer, S.C.; Lewis, J.A.; Bernhard, J.T. 3D-printed spherical dipole antenna integrated on small RF node. Electron. Lett. 2015, 51, 661. [Google Scholar] [CrossRef] [Green Version]
  133. Kimionis, J.; Isakov, M.; Koh, B.S.; Georgiadis, A.; Tentzeris, M.M. 3D-Printed Origami Packaging with Inkjet-Printed Antennas for RF Harvesting Sensors. IEEE Trans. Microw. Theory Tech. 2015, 63, 4521. [Google Scholar] [CrossRef] [Green Version]
  134. Singh, H.; Shimojima, M.; Shiratori, T.; Van An, L.; Sugamata, M.; Yang, M. Application of 3D Printing Technology in Increasing the Diagnostic Performance of Enzyme-Linked Immunosorbent Assay (ELISA) for Infectious Diseases. Sensors 2015, 15, 16503. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Guo, S.; Yang, X.; Heuzey, M.; Therriault, D. 3D printing of a multifunctional nanocomposite helical liquid sensor. Nanoscale 2015, 7, 6451. [Google Scholar] [CrossRef] [PubMed]
  136. Liao, Q.; Si, W.; Zhang, J.; Sun, H.; Qin, L. In Situ Silver Nanonets for Flexible Stretchable Electrodes. Int. J. Mol. Sci. 2023, 24, 9319. [Google Scholar] [CrossRef] [PubMed]
  137. Guo, S.; Qiu, K.; Meng, F.; Park, S.H.; McAlpine, M.C. 3D Printed Stretchable Tactile Sensors. Adv. Mater. 2017, 29, 1701218. [Google Scholar] [CrossRef]
  138. Fuh, Y.K.; Wang, B.S.; Tsai, C.-Y. Self-Powered Pressure Sensor with fully encapsulated 3D printed wavy substrate and highly-aligned piezoelectric fibers array. Sci. Rep. 2017, 7, 6759. [Google Scholar] [CrossRef] [Green Version]
  139. Bernasconi, R.; Hatami, D.; Hosseinabadi, H.N.; Zega, V.; Corigliano, A.; Suriano, R.; Levi, M.; Langfelder, G.; Magagnin, L. Hybrid additive manufacturing of a piezopolymer-based inertial sensor. Addit. Manuf. 2022, 59, 103091. [Google Scholar] [CrossRef]
  140. Raffaella, D.S. Fibre Optic Sensors for Structural Health Monitoring of Aircraft Composite Structures: Recent Advances and Applications. Sensors 2015, 15, 18666–18713. [Google Scholar] [CrossRef]
  141. Sabato, A.; Niezrecki, C.; Fortino, G. Wireless MEMS-Based Accelerometer Sensor Boards for Structural Vibration Monitoring: A Review. IEEE Sens. J. 2017, 17, 226. [Google Scholar] [CrossRef]
  142. Council, N.R. Microelectromechanical Systems: Advanced Materials and Fabrication Methods; The National Academies Press: Washington, DC, USA, 1997. [Google Scholar] [CrossRef] [Green Version]
  143. Jack, W.J. Microelectromechanical systems (MEMS): Fabrication, design and applications. Smart Mater. Struct. 2001, 10, 1115. [Google Scholar] [CrossRef] [Green Version]
  144. Hassanin, H.; Essa, K.; Elshaer, A.; Imbaby, M.; El-Mongy, H.H.; El-Sayed, T.A. Micro-fabrication of ceramics: Additive manufacturing and conventional technologies. J. Adv. Ceram. 2021, 10, 1–27. [Google Scholar] [CrossRef]
  145. de Roger, R.; Jens Ole, G.; Patrick, R.S. Fabrication and characterization of a piezoelectric accelerometer. J. Micromech. Microeng. 1999, 9, 123. [Google Scholar] [CrossRef]
  146. Khoshnoud, F.; Silva, C.W.d. Recent advances in MEMS sensor technology-mechanical applications. IEEE Instrum. Meas. Mag. 2012, 15, 14. [Google Scholar] [CrossRef]
  147. Berman, D.; Krim, J. Surface science, MEMS and NEMS: Progress and opportunities for surface science research performed on, or by, microdevices. Prog. Surf. Sci. 2013, 88, 171. [Google Scholar] [CrossRef]
  148. Hassanin, H.; Sheikholeslami, G.; Sareh, P.; Ishaq, R.B. Microadditive Manufacturing Technologies of 3D Microelectromechanical Systems. Adv. Eng. Mater. 2021, 23, 2100422. [Google Scholar] [CrossRef]
  149. Kavitha, S.; Joseph Daniel, R.; Sumangala, K. High performance MEMS accelerometers for concrete SHM applications and comparison with COTS accelerometers. Mech. Syst. Signal Process. 2016, 66–67, 410. [Google Scholar] [CrossRef]
  150. Walwadkar, S.S.; Cho, J. Evaluation of Die Stress in MEMS Packaging: Experimental and Theoretical Approaches. IEEE Trans. Compon. Packag. Technol. 2006, 29, 735. [Google Scholar] [CrossRef]
  151. Li, P.; Gao, S.; Liu, H.; Liu, J.; Shi, Y. Effects of package on the performance of MEMS piezoresistive accelerometers. Microsyst. Technol. 2013, 19, 1137. [Google Scholar] [CrossRef]
  152. Tadigadapa, S.; Mateti, K. Piezoelectric MEMS sensors: State-of-the-art and perspectives. Meas. Sci. Technol. 2009, 20, 092001. [Google Scholar] [CrossRef]
  153. Subasinghe, S.S.; Goyal, A.; Tadigadapa, S.A. High aspect ratio plasma etching of bulk Lead Zirconate Titanate. Micromach. Microfabr. Process Technol. XI 2006, 6109, 100–108. [Google Scholar] [CrossRef]
  154. Vergara, A.; Tsukamoto, T.; Fang, W.; Tanaka, S. Feedback control of thin film PZT MEMS actuator with integrated buried piezoresistors. Sens. Actuators A Phys. 2021, 332, 113131. [Google Scholar] [CrossRef]
  155. Kon, S.; Horowitz, R. A High-Resolution MEMS Piezoelectric Strain Sensor for Structural Vibration Detection. IEEE Sens. J. 2008, 8, 2027. [Google Scholar] [CrossRef]
  156. Li, C.; Jia, M.; Hong, Y.; Xue, Y.; Xiong, J. Wireless passive flexible accelerometer fabricated using micro-electro-mechanical system technology for bending structure surfaces. Front. Inf. Technol. Electron. Eng. 2022, 23, 801. [Google Scholar] [CrossRef]
  157. Gupta, S.; Giacomozzi, F.; Heidari, H.; Lorenzelli, L.; Dahiya, R. Ultra-Thin Silicon based Piezoelectric Capacitive Tactile Sensor. Procedia Eng. 2016, 168, 662. [Google Scholar] [CrossRef]
  158. Xin, Y.; Hou, T.; Liu, C.; Liu, H.; Tong, J.; Li, Y.; Lin, T. Flexible piezoelectric sensor based on PVDF-TrFE/Nanoclay composite nanofibers for physiological micro-vibration signal sensing. Measurement 2022, 201, 111742. [Google Scholar] [CrossRef]
Figure 1. Flexible piezoelectric acceleration sensors are used in behavior recognition, robotics, fault monitoring, and the Internet of Things. Reprinted with permission from Ref. [15]. Copyright 2020, iScience. Image for “robotic arms”: Reprinted with permission from Ref. [16]. Copyright 2019, John Wiley and Sons. Image for “biomedical science”: Reprinted with permission from Ref. [17]. Copyright 2017, John Wiley and Sons. Image for “wearables”: Reprinted with permission from Ref. [18]. Copyright 2018, American Chemical Society. Image for “anthropomorphic robots”: Reprinted with permission from Ref. [19]. Copyright 2018, MDPI AG. Image for “machine learning”: Reprinted with permission from Ref. [20]. Copyright 2018, Elsevier. Image for “electronic skin”: Reprinted with permission from Ref. [21]. Copyright 2018, Elsevier. Image for “aerospace”: Reprinted with permission from Ref. [22]. Copyright 2017, Elsevier. Image for “natural disasters”: Reprinted with permission from Ref. [23]. Copyright 2021, American Chemical Society. Image for “automotive industry”: Reprinted with permission from Ref. [24]. Copyright 2021, Elsevier. Image for “Smart Cities”: Reproduced under the terms of the CC-BY Creative Commons Attribution 4.0 International license (https://www.freepik.com/free-vector/isometric-smart-city-background_4358967.htm (accessed on 4 November 2022)) [25]. Image for “Smart Farming”: Reproduced under the terms of the CC-BY Creative Commons Attribution 4.0 International license (https://www.freepik.com/free-photo/smart-robotic-farmers-concept-robot-farmers-agriculture-technology-farm-automation_21544365.htm (accessed on 27 February 2023)) [26]. Image for “Smart Buildings”: Reproduced under the terms of the CC-BY Creative Commons Attribution 4.0 International license (https://www.freepik.com/free-vector/isometric-smart-city-background_4358967.htm (accessed on 4 November 2022)) [27].
Figure 1. Flexible piezoelectric acceleration sensors are used in behavior recognition, robotics, fault monitoring, and the Internet of Things. Reprinted with permission from Ref. [15]. Copyright 2020, iScience. Image for “robotic arms”: Reprinted with permission from Ref. [16]. Copyright 2019, John Wiley and Sons. Image for “biomedical science”: Reprinted with permission from Ref. [17]. Copyright 2017, John Wiley and Sons. Image for “wearables”: Reprinted with permission from Ref. [18]. Copyright 2018, American Chemical Society. Image for “anthropomorphic robots”: Reprinted with permission from Ref. [19]. Copyright 2018, MDPI AG. Image for “machine learning”: Reprinted with permission from Ref. [20]. Copyright 2018, Elsevier. Image for “electronic skin”: Reprinted with permission from Ref. [21]. Copyright 2018, Elsevier. Image for “aerospace”: Reprinted with permission from Ref. [22]. Copyright 2017, Elsevier. Image for “natural disasters”: Reprinted with permission from Ref. [23]. Copyright 2021, American Chemical Society. Image for “automotive industry”: Reprinted with permission from Ref. [24]. Copyright 2021, Elsevier. Image for “Smart Cities”: Reproduced under the terms of the CC-BY Creative Commons Attribution 4.0 International license (https://www.freepik.com/free-vector/isometric-smart-city-background_4358967.htm (accessed on 4 November 2022)) [25]. Image for “Smart Farming”: Reproduced under the terms of the CC-BY Creative Commons Attribution 4.0 International license (https://www.freepik.com/free-photo/smart-robotic-farmers-concept-robot-farmers-agriculture-technology-farm-automation_21544365.htm (accessed on 27 February 2023)) [26]. Image for “Smart Buildings”: Reproduced under the terms of the CC-BY Creative Commons Attribution 4.0 International license (https://www.freepik.com/free-vector/isometric-smart-city-background_4358967.htm (accessed on 4 November 2022)) [27].
Coatings 13 01252 g001
Figure 2. Development and common structures of flexible piezoelectric accelerometers. The development of flexible piezoelectric accelerometers can be divided into three parts: (a) General accelerometers. Reprinted with permission from Ref. [28]. Copyright 2018, Elsevier. (b) Flexible accelerometers. Reprinted with permission from Ref. [29]. Copyright 2020, Elsevier. (c) Flexible piezoelectric accelerometers. Reprinted with permission from Ref. [30]. Copyright 2017, Elsevier. Common structures for flexible piezoelectric accelerometers: (d,e) The cantilever beam structure. Reprinted with permission from Ref. [31]. Copyright 2018, Taylor & Francis and The Author(s). (f,g) Tri-axis structure. Reprinted with permission from Ref. [32]. Copyright 2016, Science Advances. (h,i) Sandwich structure. Reprinted with permission from Ref. [33]. Copyright 2015, John Wiley and Sons. Reprinted with permission from Ref. [34]. Copyright 2021, Elsevier. (j,k) Micropillar array. Reprinted with permission from Ref. [35]. Copyright 2017, John Wiley and Sons.
Figure 2. Development and common structures of flexible piezoelectric accelerometers. The development of flexible piezoelectric accelerometers can be divided into three parts: (a) General accelerometers. Reprinted with permission from Ref. [28]. Copyright 2018, Elsevier. (b) Flexible accelerometers. Reprinted with permission from Ref. [29]. Copyright 2020, Elsevier. (c) Flexible piezoelectric accelerometers. Reprinted with permission from Ref. [30]. Copyright 2017, Elsevier. Common structures for flexible piezoelectric accelerometers: (d,e) The cantilever beam structure. Reprinted with permission from Ref. [31]. Copyright 2018, Taylor & Francis and The Author(s). (f,g) Tri-axis structure. Reprinted with permission from Ref. [32]. Copyright 2016, Science Advances. (h,i) Sandwich structure. Reprinted with permission from Ref. [33]. Copyright 2015, John Wiley and Sons. Reprinted with permission from Ref. [34]. Copyright 2021, Elsevier. (j,k) Micropillar array. Reprinted with permission from Ref. [35]. Copyright 2017, John Wiley and Sons.
Coatings 13 01252 g002
Figure 3. Relationships between piezoelectric acceleration sensors’ parameters. (a) The waveform of output voltage. Reprinted with permission from Ref. [44]. Copyright 2016, Elsevier. (b) The relationship between sensitivity and frequency. Reprinted with permission from Ref. [45]. Copyright 2005, Elsevier. (c) The relationship between sensitivity and pre-tightening torque. Reprinted with permission from Ref. [46]. Copyright 2012, Elsevier. (d) Sensitivity as a function of temperature. Reprinted with permission from Ref. [48]. Copyright 2010, AIP Publishing. (e) Noise characteristics of piezoelectric acceleration sensors. Reprinted with permission from Ref. [50]. Copyright 2000, Elsevier. (f) Output charge as a function of acceleration. Reprinted with permission from Ref. [48]. Copyright 2010, AIP Publishing.
Figure 3. Relationships between piezoelectric acceleration sensors’ parameters. (a) The waveform of output voltage. Reprinted with permission from Ref. [44]. Copyright 2016, Elsevier. (b) The relationship between sensitivity and frequency. Reprinted with permission from Ref. [45]. Copyright 2005, Elsevier. (c) The relationship between sensitivity and pre-tightening torque. Reprinted with permission from Ref. [46]. Copyright 2012, Elsevier. (d) Sensitivity as a function of temperature. Reprinted with permission from Ref. [48]. Copyright 2010, AIP Publishing. (e) Noise characteristics of piezoelectric acceleration sensors. Reprinted with permission from Ref. [50]. Copyright 2000, Elsevier. (f) Output charge as a function of acceleration. Reprinted with permission from Ref. [48]. Copyright 2010, AIP Publishing.
Coatings 13 01252 g003
Figure 4. (ac) Atomic structures, SEM images, and related components of ZnO, respectively. Reprinted with permission from Ref. [64]. Copyright 2019, Elsevier. Reprinted with permission from Ref. [65]. Copyright 2009, American Chemical Society. Reprinted with permission from Ref. [66]. Copyright 2018, American Chemical Society. (df) Atomic structures, SEM images, and related components of PZT, respectively. Reprinted with permission from Ref. [68]. Copyright 2012, Elsevier. Reprinted with permission from Ref. [68]. Copyright 2015, Elsevier. Reprinted with permission from Ref. [69]. Copyright 2019, Elsevier. (gi) Atomic structures, SEM images, and related components of BaTiO3, respectively. Reprinted with permission from Ref. [70]. Copyright 2022, John Wiley and Sons. Reprinted with permission from Ref. [71]. Copyright 2011, American Chemical Society. Reprinted with permission from Ref. [72]. Copyright 2014, Elsevier. (jl) Atomic structures, SEM images, and related components of PVDF, respectively. Reprinted with permission from Ref. [73]. Copyright 2012, Elsevier. Reprinted with permission from Ref. [74]. Copyright 2014, Elsevier. Reprinted with permission from Ref. [75]. Copyright 2020, MDPI AG.
Figure 4. (ac) Atomic structures, SEM images, and related components of ZnO, respectively. Reprinted with permission from Ref. [64]. Copyright 2019, Elsevier. Reprinted with permission from Ref. [65]. Copyright 2009, American Chemical Society. Reprinted with permission from Ref. [66]. Copyright 2018, American Chemical Society. (df) Atomic structures, SEM images, and related components of PZT, respectively. Reprinted with permission from Ref. [68]. Copyright 2012, Elsevier. Reprinted with permission from Ref. [68]. Copyright 2015, Elsevier. Reprinted with permission from Ref. [69]. Copyright 2019, Elsevier. (gi) Atomic structures, SEM images, and related components of BaTiO3, respectively. Reprinted with permission from Ref. [70]. Copyright 2022, John Wiley and Sons. Reprinted with permission from Ref. [71]. Copyright 2011, American Chemical Society. Reprinted with permission from Ref. [72]. Copyright 2014, Elsevier. (jl) Atomic structures, SEM images, and related components of PVDF, respectively. Reprinted with permission from Ref. [73]. Copyright 2012, Elsevier. Reprinted with permission from Ref. [74]. Copyright 2014, Elsevier. Reprinted with permission from Ref. [75]. Copyright 2020, MDPI AG.
Coatings 13 01252 g004
Figure 5. Materials for flexible piezoelectric acceleration sensors: (a) Microscopic morphology of PVDF-TERE nanofiber. Reprinted with permission from Ref. [77]. Copyright 2018, Elsevier. (bd) The process of piezoelectric ceramics is compounded with piezoelectric organic materials. Reprinted with permission from Ref. [83]. Copyright 2014, John Wiley and Sons. Reprinted with permission from Ref. [84]. Copyright 2020, Elsevier. Reprinted with permission from Ref. [85]. Copyright 2019, Hindawi. (e,f) Composite process of piezoelectric ceramics and non-piezoelectric organic materials and its high flexibility (g) and high rolling capability (h). Reprinted with permission from Ref. [86]. Copyright 2017, American Chemical Society.
Figure 5. Materials for flexible piezoelectric acceleration sensors: (a) Microscopic morphology of PVDF-TERE nanofiber. Reprinted with permission from Ref. [77]. Copyright 2018, Elsevier. (bd) The process of piezoelectric ceramics is compounded with piezoelectric organic materials. Reprinted with permission from Ref. [83]. Copyright 2014, John Wiley and Sons. Reprinted with permission from Ref. [84]. Copyright 2020, Elsevier. Reprinted with permission from Ref. [85]. Copyright 2019, Hindawi. (e,f) Composite process of piezoelectric ceramics and non-piezoelectric organic materials and its high flexibility (g) and high rolling capability (h). Reprinted with permission from Ref. [86]. Copyright 2017, American Chemical Society.
Coatings 13 01252 g005
Figure 6. Structure of flexible piezoelectric acceleration sensor. (a) Structure of a conventional piezoelectric acceleration sensor. Reprinted with permission from Ref. [83]. Copyright 2014, John Wiley and Sons. (b,c) The design and testing of a uniaxial piezoelectric acceleration sensor. Reprinted with permission from Ref. [92]. Copyright 2018, MDPI AG. (d) Schematic illustration of the flexible hNCG composed of BT NW-embedded P(VDF-TrFE). Reprinted with permission from Ref. [93]. Copyright 2018, John Wiley and Sons. (e) Schematic diagram of the structure of a low-detection-limit bionic piezoelectric sensor. Reprinted with permission from Ref. [94]. Copyright 2022, Elsevier. (f) A self-powered, high-sensitivity acceleration sensor based on a friction-electric nanogenerator design. Reprinted with permission from Ref. [95]. Copyright 2017, American Chemical Society. (g) The fabrication process of a flexible piezoelectric sensor made of group III nitride film. Reprinted with permission from Ref. [96]. Copyright 2019, Elsevier. (h) A sensor patch that detects heartbeat and respiration signals. Reprinted with permission from Ref. [97]. Copyright 2013, Elsevier.
Figure 6. Structure of flexible piezoelectric acceleration sensor. (a) Structure of a conventional piezoelectric acceleration sensor. Reprinted with permission from Ref. [83]. Copyright 2014, John Wiley and Sons. (b,c) The design and testing of a uniaxial piezoelectric acceleration sensor. Reprinted with permission from Ref. [92]. Copyright 2018, MDPI AG. (d) Schematic illustration of the flexible hNCG composed of BT NW-embedded P(VDF-TrFE). Reprinted with permission from Ref. [93]. Copyright 2018, John Wiley and Sons. (e) Schematic diagram of the structure of a low-detection-limit bionic piezoelectric sensor. Reprinted with permission from Ref. [94]. Copyright 2022, Elsevier. (f) A self-powered, high-sensitivity acceleration sensor based on a friction-electric nanogenerator design. Reprinted with permission from Ref. [95]. Copyright 2017, American Chemical Society. (g) The fabrication process of a flexible piezoelectric sensor made of group III nitride film. Reprinted with permission from Ref. [96]. Copyright 2019, Elsevier. (h) A sensor patch that detects heartbeat and respiration signals. Reprinted with permission from Ref. [97]. Copyright 2013, Elsevier.
Coatings 13 01252 g006
Figure 7. (a) Schematic diagram of the overall process of 3D printing. Reprinted with permission from Ref. [101]. Copyright 2017, MDPI AG. (b) 3D printing of flex sensors: (i) the CAD design of the flex sensor, (ii) the printed flex sensor, (iii) the printed sensor undergoing flexing; (c) 3D printing of a “glove”: (i) CAD design of the 3D-printed “glove”, (ii) the printed “glove”, (iii) the printed “glove” before flexing, (iv) the printed “glove” during flexing. Reprinted with permission from Ref. [102]. Copyright 2012, Plos ONE. (dh) Custom PuSL fabrication system for piezoelectric nanocomposite and 3D-printed piezoelectric complex structures with fine surface finish. Reprinted with permission from Ref. [103]. Copyright 2019, John Wiley and Sons. (il) 1D, 2D, 2.5D, 3D shapes printed with PVDF. Reprinted with permission from Ref. [104]. Copyright 2017, American Chemical Society.
Figure 7. (a) Schematic diagram of the overall process of 3D printing. Reprinted with permission from Ref. [101]. Copyright 2017, MDPI AG. (b) 3D printing of flex sensors: (i) the CAD design of the flex sensor, (ii) the printed flex sensor, (iii) the printed sensor undergoing flexing; (c) 3D printing of a “glove”: (i) CAD design of the 3D-printed “glove”, (ii) the printed “glove”, (iii) the printed “glove” before flexing, (iv) the printed “glove” during flexing. Reprinted with permission from Ref. [102]. Copyright 2012, Plos ONE. (dh) Custom PuSL fabrication system for piezoelectric nanocomposite and 3D-printed piezoelectric complex structures with fine surface finish. Reprinted with permission from Ref. [103]. Copyright 2019, John Wiley and Sons. (il) 1D, 2D, 2.5D, 3D shapes printed with PVDF. Reprinted with permission from Ref. [104]. Copyright 2017, American Chemical Society.
Coatings 13 01252 g007
Figure 9. Applications and fabrication processes of 3D-printed flexible piezoelectric acceleration sensor. (a) A three-view drawing of a flexible tactile sensor and its fabrication processes is shown in (b). Reprinted with permission from Ref. [137]. Copyright 2017, John Wiley and Sons. (c) A piezoelectric pressure sensor manufactured using 3D printing. Reprinted with permission from Ref. [138]. Copyright 2017, Springer Nature. (d) Photograph of a piezopolymer-based inertial sensor that uses hybrid additive manufacturing methods. Reprinted with permission from Ref. [139]. Copyright 2022, Elsevier.
Figure 9. Applications and fabrication processes of 3D-printed flexible piezoelectric acceleration sensor. (a) A three-view drawing of a flexible tactile sensor and its fabrication processes is shown in (b). Reprinted with permission from Ref. [137]. Copyright 2017, John Wiley and Sons. (c) A piezoelectric pressure sensor manufactured using 3D printing. Reprinted with permission from Ref. [138]. Copyright 2017, Springer Nature. (d) Photograph of a piezopolymer-based inertial sensor that uses hybrid additive manufacturing methods. Reprinted with permission from Ref. [139]. Copyright 2022, Elsevier.
Coatings 13 01252 g009
Figure 10. Comparison of three fabrication processes for piezoelectric MEMS sensors.
Figure 10. Comparison of three fabrication processes for piezoelectric MEMS sensors.
Coatings 13 01252 g010
Figure 12. Application of MEMS flexible piezoelectric acceleration sensors. (a) Fabrication process and structure of a wireless passive flexible accelerometer fabricated using micro-electro-mechanical system technology for bending structure surfaces. (b) Steps to fabricate ultra-thin silicon-based PVDF-TrFE capacitors. (c) (top) Scanning electron microscopy image of the piezo-capacitor sensor; (bottom) optical profilometer image showing warp image of a thin chip with the piezo-capacitor. Reprinted with permission from Ref. [157]. Copyright 2016, Elsevier. (d,e) In situ assembly of PMVS sensor. (f) The relationship between PMVS sensors’ sensitivity and nanoclay content. Reprinted with permission from Ref. [158]. Copyright 2022, Elsevier.
Figure 12. Application of MEMS flexible piezoelectric acceleration sensors. (a) Fabrication process and structure of a wireless passive flexible accelerometer fabricated using micro-electro-mechanical system technology for bending structure surfaces. (b) Steps to fabricate ultra-thin silicon-based PVDF-TrFE capacitors. (c) (top) Scanning electron microscopy image of the piezo-capacitor sensor; (bottom) optical profilometer image showing warp image of a thin chip with the piezo-capacitor. Reprinted with permission from Ref. [157]. Copyright 2016, Elsevier. (d,e) In situ assembly of PMVS sensor. (f) The relationship between PMVS sensors’ sensitivity and nanoclay content. Reprinted with permission from Ref. [158]. Copyright 2022, Elsevier.
Coatings 13 01252 g012
Table 1. Performance comparison of different types of piezoelectric acceleration sensors.
Table 1. Performance comparison of different types of piezoelectric acceleration sensors.
TypesMold Area (mm × mm)Frequency Range (kHz)SensitivityReferences
Resistive type 15 × 513.30.545 mV/g[51]
Resistive type 23.5 × 3.51.54 mV/g[52]
Capacitive type6.5 × 56/[53]
PZT6 × 617.414 mV/g[54]
AlN6.8 × 6.81.1400 mV/g[55]
AlN_C5002.3 × 2.312.5355 mV/g[50]
AlN_S2002.3 × 2.342.8176 mV/g[50]
YCOB/0.62.4 pC/g[48]
BTS/32.6 pC/g[47]
N/A/63.5 pC/g[56]
Table 2. Three-dimensional printing technology’s classification, commonly used materials, advantages, and disadvantages.
Table 2. Three-dimensional printing technology’s classification, commonly used materials, advantages, and disadvantages.
CategoriesMaterialsAdvantagesDisadvantages
FDMAll-plastic materials
(PA, PLA, PC, etc.)
Low cost
Open-source code
Slow printing speed
Low resolution
DIWCeramics, plastics, hydrogels, and even living cellsFast printing speed
High durability
Suitable for repairing components
Subsequent processing required
Low mechanical properties
Poor surface finish
SLA, DLPResins, polymers, plasticsHigh speed
High resolution
Complex and high-precision structure
Single material
Easy to damage the device
Support structure required
LOMPapers, plastics, metalsLow Cost
High speed
Color printing
Post-processing required
Low resolution
Design limitations
SLS, SLMPlastics, metals, ceramics, waxesHigh strength and stiffness
Suitable for aerospace and military
No support mechanism required
Slow printing speed
Size limitation
High power usage
PloyjetPhotosensitive resinsHigh resolution, suitable for small and delicate devicesLow strength
3DPCeramics, plasters, sugarsCan print simultaneously
Color printing
No support mechanism required
Poor surface strength and roughness
Requires post treatment
Low mechanical properties
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Liao, Y.; Yang, H.; Liao, Q.; Si, W.; Chu, Y.; Chu, X.; Qin, L. A Review of Flexible Acceleration Sensors Based on Piezoelectric Materials: Performance Characterization, Parametric Analysis, Frontier Technologies, and Applications. Coatings 2023, 13, 1252. https://doi.org/10.3390/coatings13071252

AMA Style

Liao Y, Yang H, Liao Q, Si W, Chu Y, Chu X, Qin L. A Review of Flexible Acceleration Sensors Based on Piezoelectric Materials: Performance Characterization, Parametric Analysis, Frontier Technologies, and Applications. Coatings. 2023; 13(7):1252. https://doi.org/10.3390/coatings13071252

Chicago/Turabian Style

Liao, Yaoyao, Hong Yang, Qingwei Liao, Wei Si, Yu Chu, Xiangcheng Chu, and Lei Qin. 2023. "A Review of Flexible Acceleration Sensors Based on Piezoelectric Materials: Performance Characterization, Parametric Analysis, Frontier Technologies, and Applications" Coatings 13, no. 7: 1252. https://doi.org/10.3390/coatings13071252

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop